nucleoside phosphorylases from thermophiles

152
Nucleoside phosphorylases from thermophiles Recombinant expression and biocatalytic use for modified nucleosides Kathleen Szeker, Berlin 2012

Upload: others

Post on 12-Jan-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Nucleoside phosphorylases from thermophiles

Recombinant expression and biocatalytic use for modified nucleosides

Kathleen Szeker, Berlin 2012

Nucleoside phosphorylases from thermophiles

Recombinant expression and biocatalytic use for modified nucleosides

vorgelegt von

Diplom-Ingenieurin Kathleen Szeker

aus Potsdam

von der Fakultät III – Prozesswissenschaften

der Technischen Universität Berlin

zur Erlangung des akademischen Grades

Doktor der Ingenieurwissenschaften

- Dr.-Ing. –

genehmigte Dissertation

Promotionsausschuss:

Vorsitzender Prof. Dr. Roland Lauster

Gutachter Prof. Dr. Peter Neubauer

Gutachterin Prof. Dr. Marion Ansorge-Schumacher

Gutachter Prof. Dr. Igor A. Mikhailopulo

Tag der wissenschaftlichen Aussprache: 11.06.2012

Berlin 2012

D83

Abstract

Modified nucleosides are valuable pharmaceutical agents used in the treatment of cancer and viral

infections. Moreover, they serve as building blocks in the synthesis of therapeutic oligonucleotides

with advanced properties.

While the chemical modification of pyrimidine nucleosides is generally well established, the synthesis

of modified purine nucleosides is often rather challenging, resulting in multistage processes with low

yield. Alternative synthetic routes include the chemo-enzymatic synthesis of purine nucleosides from

a pyrimidine nucleoside serving as pentofuranosyl donor and a purine base functioning as

pentofuranosyl acceptor. As biocatalysts, nucleoside phosphorylases (NPs) are used to catalyze the

regio- and stereoselective transfer reaction, whereby natural or chemically prepared artificial

precursors can be applied as substrate. Unfortunately, a number of highly interesting nucleoside

analogues are hardly recognized as substrate by NPs that are currently in use. Moreover, high

temperatures are desirable to increase the concentration of poorly soluble purine bases, but many

enzymes are rapidly deactivated by heat. Both factors limit the scope and the efficiency of NP

mediated syntheses of modified nucleosides and prompted us to study novel, thermostable

nucleoside phosphorylase variants as potential biocatalysts.

Therefore a set of 5 NPs from 4 different thermophilic microorganisms (Deinococcus geothermalis,

Geobacillus thermoglucosidasius, Thermus thermophilus, Aeropyrum pernix) has been overexpressed

in E. coli. The recombinant proteins were characterized in order to assess their potential application

as biocatalysts. Thermal properties (temperature optima, stability) varied significantly and were

dependent on the source microorganism and the type of enzyme. Investigations of the substrate

specificities revealed striking differences in the ability to tolerate modified nucleosides as substrate.

The data allowed us to select and test the most promising combinations of enzymes for enzymatic

transglycosylation reactions. In focus of the present work was thereby the synthesis of 2′-fluorinated

purine nucleosides as well as 2,6-dihalogenated purine nucleosides. 2′-Fluorinated nucleosides were

found to have valuable pharmaceutical properties and impart favourable characteristics to synthetic

oligonucleotides. On the other hand, 2,6-dihalogenated purine nucleosides are versatile precursors

for a variety of purine modified nucleosides. In comparison to E. coli enzymes that are described in

literature as biocatalysts for the synthesis of 2′-fluorinated purine nucleosides, the application of the

novel, thermostable enzymes permits the operation at higher temperature, and appears to be more

efficient in the synthesis 2′-fluorinated purine nucleosides. Furthermore, 2,6-dihalogenated purines

were readily accepted as substrates and the respective (deoxy-)ribosides were rapidly produced by

the novel enzyme preparations.

The results corroborate the general potential of thermostable NPs in the synthesis of modified

nucleosides and specifically pave the way towards improved, environmentally friendly synthetic

procedures affording valuable 2′-fluorinated and 2,6-dihalogenated purine nucleoside analogues.

The present work was performed from April 2009 – April 2012 in the research group of Prof. Dr.

Peter Neubauer (Laboratory of Bioprocess Engineering) at the Department of Biotechnology,

Technische Universität Berlin.

Publications

Szeker, K., Niemitalo, O., Casteleijn, M.G., Juffer, A.H., and Neubauer, P., 2011.

“High-temperature cultivation and 5' mRNA optimization are key factors for the efficient overexpression of thermostable Deinococcus geothermalis purine nucleoside phosphorylase in Escherichia coli”

Journal of Biotechnology, 156(4), 268-274

Szeker, K., Zhou, X., Schwab, T., Casanueva, A., Cowan, D., Mikhailopulo, I.A., and Neubauer, P., 2012.

“Comparative investigations on thermostable pyrimidine nucleoside phosphorylases from Geobacillus thermoglucosidasius and Thermus thermophilus.”

Journal of Molecular Catalysis B: Enzymatic, in press

Conference contributions

K. Szeker, X. Zhou, A. Scholz, M. Ansorge-Schumacher, I. A. Mikhailopulo, and P. Neubauer

“Thermostable nucleoside phosphorylases for the synthesis of purine nucleoside analogues”

X. Zhou, K. Szeker, I. A. Mikhailopulo, and P. Neubauer

“Thermostable biocatalysts with purine nucleoside activity”

Catalyzing Bio-Economy – Biocatalysts for industrial biotechnology, Annual meeting of the DECHEMA-VAAM-Section Biotransformations, Frankfurt, Germany, April 2012

K. Szeker, X. Zhou, I. A. Mikhailopulo, and P. Neubauer

“Characterizing thermostable nucleoside phosphorylases for their use as biocatalysts”

Biotrans 2011, Giardini Naxos, Italy, October 2011

K. Szeker, X. Zhou, I. A. Mikhailopulo, and P. Neubauer

“Overexpression and biocatalytic characterization of thermostable nucleoside phosphorylases in Escherichia coli”

1st European Congress of Applied Biotechnology, Berlin, Germany, September 2011

K. Szeker, O. Niemitalo, and P. Neubauer

“Increasing the expression level of purine nucleoside phosphorylase from thermophilic origin in Escherichia coli”

6th Conference on recombinant protein production – A comparative view on host physiology, Vienna, Austria, February 2011

K. Szeker, M. Casteleijn, and P. Neubauer

“Optimization of soluble expression of recombinant thermophilic nucleoside phosphorylases”

14th International Biotechnology Symposium and Exhibition, Rimini, Italy, September 2010

K. Szeker, and P. Neubauer

“Novel biocatalysts for the preparation of modified nucleosides”

Talk at BIG-NSE / UniCat Mini Symposium – Protein Engineering, Berlin, Germany, May 2010

K. Szeker, and P. Neubauer

“Biocatalysts in nucleoside chemistry”

Talk at German-Hungarian biotechnology seminar, Miskolc, Hungary, July 2009

List of abbreviations

AdoP Adenosine phosphorylase

A. hydrophila Aeromonas hydrophila

A. pernix Aeropyrum pernix

AhPNP Purine nucleoside phosphorylase from A. hydrophila

Anhydro-Urd O2,2′-Anhydro-1-(-D-arabinofuranosyl)uracil

ApMTAP 5′-Methylthioadenosine phosphorylase from A. pernix

ApUP Uridine phosphorylase of A. pernix

ara-A 9-(-D-Arabinofuranosyl)adenine

ara-U 1-(-D-Arabinofuranosyl)uracil

ASOs Antisense oligonucleotides

AZT Azidothymidine

B. cereus Bacillus cereus

B. subtilis Bacillus subtilis

CAI Codon adaption index

D. geothermalis Deinococcus geothermalis

DgPNP Purine nucleoside phosphorylase from D. geothermalis

dAdo2′F 9-(2-Deoxy-2-fluoro-β-D-arabinofuranosyl)adenine

dAdo2′F 2′-Deoxy-2′-fluoroadenosine

dUrd2′F 1-(2-Deoxy-2-fluoro-β-D-arabinofuranosyl)uracil

dUrd2′F 2′-Deoxy-2′-fluorouridine

DTT Dithiothreitol

EcPNP Purine nucleoside phosphorylase of E. coli

EcTP Thymidine phosphorylase of E. coli

EcUP Uridine phosphorylase of E. coli

GsPyNP Pyrimidine nucleoside phosphorylase from G. stearothermophilus

G. stearothermophilus Geobacillus stearothermophilus

G. thermoglucosidasius Geobacillus thermoglucosidasius

GtPNP Purine nucleoside phosphorylase from G. thermoglucosidasius

GtPyNP Pyrimidine nucleoside phosphorylase from G. thermoglucosidasius

HPLC High-performance liquid chromatography

IPTG Isopropyl β-D-1-thiogalactopyranoside

KP buffer Potassium phosphate buffer

LB medium Lysogeny broth medium

MTAP 5′-Methylthioadenosine phosphorylase

NdRT N-deoxyribosyltransferase

NP Nucleoside phosphorylase

NP buffer Sodium phosphate buffer

PNP Purine nucleoside phosphorylase

PyNP Pyrimidine nucleoside phosphorylase

P. furiosus Pyrococcus furiosus

S. solfataricus Sulfolobus solfataricus

SDS-PAGE Sodium dodecyl sulfate polyacrylamide gel electrophoresis

TB medium Terrific broth medium

T. thermophilus Thermus thermophilus

TP Thymidine phosphorylase

TtPyNP Thymidine phosphorylase from T. thermophilus

UP Uridine phosphorylase

XanoP Xanthosine phosphorylase

Table of contents

1. Introduction ............................................................................................. 5

1.1. Modified nucleosides in pharmacy and life sciences ............................................. 5

1.1.1. Natural nucleosides .............................................................................................. 5

1.1.2. Modified nucleosides in clinical use ..................................................................... 5

1.1.3. Modified nucleosides as building blocks for synthetic nucleic acids ................... 9

1.2. Synthetic routes towards nucleoside analogues .................................................. 11

1.2.1. Chemical synthesis ............................................................................................. 11

1.2.2. Enzymes for nucleoside synthesis ...................................................................... 12

1.2.3. The NP-catalyzed transglycosylation of nucleosides ......................................... 14

1.2.4. Formulation of NPs for technical use ................................................................. 15

1.3. Nucleoside phosphorylases ................................................................................. 16

1.3.1. Physiological role ................................................................................................ 16

1.3.2. Basic nature of the catalytic mechanism ........................................................... 17

1.3.3. Classification of NPs ........................................................................................... 20

1.3.4. From natural to recombinant production of nucleoside phosphorylases ......... 22

1.4. Motivation and structure of the present work .................................................... 24

1.4.1. Enzyme-assisted synthesis of modified purine nucleosides .............................. 24

1.4.2. Taking advantage of thermostable nucleoside phosphorylases ........................ 25

1.4.3. Recombinant expression of thermostable NPs in E. coli ................................... 25

1.4.4. Experimental outline and key objectives ........................................................... 25

2. Experimental part ................................................................................... 27

2.1. Generation of expression plasmids ..................................................................... 27

2.1.1. Gene isolation .................................................................................................... 27

2.1.2. Plasmids .............................................................................................................. 27

2.1.3. Recombinational cloning .................................................................................... 28

2.1.4. Cloning via restriction and digestion .................................................................. 29

2.1.5. Verification of the cloning steps and propagation of vector constructs ........... 30

2.1.6. Site-directed mutagenesis .................................................................................. 30

2 Introduction

2.2. Bioinformatics ..................................................................................................... 32

2.2.1. Amino acid sequence analysis and homology modelling .................................. 32

2.2.2. Secondary mRNA prediction and sequence optimization ................................. 32

2.3. Bacterial growth and recombinant protein expression ........................................ 33

2.3.1. Preparation of recombinant E. coli cell banks ................................................... 33

2.3.2. Recombinant protein expression ....................................................................... 33

2.4. Preparation of protein samples ........................................................................... 34

2.4.1. Cell disruption .................................................................................................... 34

2.4.2. Protein purification ............................................................................................ 34

2.5. Protein analytics ................................................................................................. 35

2.5.1. SDS-PAGE analysis .............................................................................................. 35

2.5.2. Determination of the protein concentration ..................................................... 35

2.5.3. Protein unfolding studies ................................................................................... 36

2.6. Activity assays ..................................................................................................... 36

2.6.1. Spectroscopic assay for PNP activity .................................................................. 36

2.6.2. Standard assay with purified proteins ............................................................... 37

2.6.3. Thermal properties of the enzymes ................................................................... 37

2.6.4. Kinetic parameters ............................................................................................. 38

2.6.5. Substrate screenings .......................................................................................... 38

2.6.6. Synthetic reactions ............................................................................................. 38

2.7. HPLC analysis ...................................................................................................... 38

3. Recombinant expression of nucleoside phosphorylases .......................... 41

3.1. Introduction ........................................................................................................ 41

3.1.1. Recombinant expression of thermostable proteins in E. coli ............................ 41

3.1.2. Target enzymes of this study ............................................................................. 42

3.2. Sequence analysis and theoretical predictions .................................................... 44

3.3. Expression of DgPNP ........................................................................................... 45

3.3.1. Towards the functional expression of DgPNP .................................................... 45

3.3.2. DgPNP expression optimization by reducing secondary 5′mRNA stability ........ 48

Introduction 3

3.3.3. Functional expression of DgPNP with N-terminal hexahistidine tag ................. 53

3.3.4. DgPNP expression - summary and conclusions ................................................. 55

3.4. Expression of ApMTAP ........................................................................................ 56

3.4.1. Expression of the wild type ApMTAP gene without tag .................................... 56

3.4.2. ApMTAP expression with N-terminal hexahistidine tag .................................... 57

3.4.3. ApMTAP - summary and conclusions ................................................................. 62

3.5. Expression of GtPNP ........................................................................................... 62

3.5.1. GtPNP expression with C-terminal hexahistidine tag ........................................ 63

3.5.2. GtPNP expression with N-terminal hexahistidine tag ........................................ 63

3.5.3. GtPNP expression - summary and conclusion ................................................... 63

3.6. Expression of GtPyNP .......................................................................................... 64

3.6.1. Chemical lysis buffer decreases apparent thermal stability of GtPyNP ............. 64

3.6.2. GtPyNP expression with N-terminal hexahistidine tag ...................................... 65

3.6.3. GtPyNP expression - summary and conclusions ................................................ 66

3.7. Expression of TtPyNP .......................................................................................... 66

3.7.1. TtPyNP expression with N-terminal hexahistidine tag ...................................... 66

3.7.2. TtPyNP expression - summary and conclusions ................................................. 68

3.8. Expression of ApUP ............................................................................................. 68

3.8.1. Expression of ApUP without tag ......................................................................... 69

3.8.2. Expression of ApUP with N-terminal hexahistidine tag ..................................... 71

3.8.3. ApUP expression - summary and discussion ...................................................... 71

3.9. Recombinant expression of NPs - summary and conclusions ............................... 72

4. Characterization of thermostable NPs ..................................................... 75

4.1. Thermostable PyNPs ........................................................................................... 75

4.1.1. Homology modelling .......................................................................................... 76

4.1.2. Thermal characteristics ...................................................................................... 76

4.1.3. Kinetic parameters ............................................................................................. 78

4.1.4. Phosphorolysis of 2′-fluorosubstituted pyrimidine nucleosides ........................ 80

4.1.5. PNP activity of PyNPs ......................................................................................... 83

4 Introduction

4.1.6. Characterization of thermostable PyNPs - summary and conclusions .............. 84

4.2. Thermostable enzymes with PNP activity ........................................................... 85

4.2.1. Sequence analysis and homology modelling ..................................................... 86

4.2.2. Thermal characteristics ...................................................................................... 88

4.2.3. Substrate specificities ......................................................................................... 91

4.2.4. Characterization of thermostable PNP enzymes - summary and conclusions .. 95

5. Enzymatic transglycosylations with thermostable NPs ............................ 97

5.1. Introduction ........................................................................................................ 97

5.1.1. Chemical synthesis of 2′-fluorinated nucleosides .............................................. 97

5.1.2. The chemo-enzymatic synthesis of 2′-fluorinated purine nucleosides ............. 99

5.1.3. 2,6-Dihalogenated purine nucleosides ............................................................ 100

5.2. Synthesis of 2′-fluorosubstituted purine nucleosides ......................................... 101

5.2.1. Synthesis of 2′-deoxy-2′-fluoroadenosine ........................................................ 101

5.2.2. Synthesis of 9-(2-deoxy-2-fluoro--D-arabinofuranosyl)adenine ................... 104

5.3. Synthesis of 2,6-dihalogenated purine nucleosides ............................................ 107

5.3.1. Synthesis of 2,6-dihalogenated purine ribosides ............................................. 108

5.3.2. Synthesis of 2,6-dihalogenated purine deoxyribosides ................................... 110

5.4. Enzymatic transglycosylations– summary and conclusions ................................ 112

5.4.1. 2′-Fluorinated purine nucleosides .................................................................... 112

5.4.2. 2,6-Dihalogenated purine nucleosides ............................................................ 113

6. Final conclusions ................................................................................... 115

References .................................................................................................. 119

Appendix ..................................................................................................... 135

Zusammenfassung ...................................................................................... 137

Acknowledgements ..................................................................................... 138

Introduction 5

1. Introduction

Nucleoside phosphorylases (NPs) are enzymes that catalyse the reversible phosphorolysis of

nucleosides. The purpose of the present study is to utilize thermostable variants for the chemo-

enzymatic synthesis of modified nucleosides. The major focus of this chapter is to give an overview

about the underlying motivation. First, the pharmaceutical and scientific value of nucleoside

analogues will be demonstrated. Next, chemical and chemo-enzymatic routes towards their

preparation will be discussed, followed by a closer glance on NPs. Finally the approach of this work

will be explained and the experimental strategy presented.

1.1. Modified nucleosides in pharmacy and life sciences

Natural nucleosides are major constituents of nucleic acids. Since they play a key role for the cellular

life, reflected by the storage of genetic information or metabolic regulation, it is obvious that

modifications can turn them into highly bioactive compounds. Therefore modified nucleosides find

broad application in both pharmaceutical industry and as molecular biological tools.

1.1.1. Natural nucleosides

Nucleosides consist out of a nucleobase that is connected to the anomeric centre (C1′) of a sugar via

a β-glycosidic linkage. In natural nucleosides this sugar represents either a D-ribose or a

D-2-deoxyribose. Depending on the heterocyclic base, nucleosides are classified as purine or

pyrimidine nucleosides. While purine bases are linked via the N-9 atom, pyrimidine bases are linked

via the N1 atom to the sugar moiety. In Figure 1 the four natural deoxynucleosides that constitute

the building blocks of our DNA are shown.

O

OH

OH

NH

N

O

O

Thymidine

O

OH

OH

N

NN

N

NH2

Deoxyadenosine

O

OH

OH

NH

NN

N

O

Deoxyguanosine

NH2

CH3

O

OH

OH

N

N

NH2

O

Deoxycytidine

1'2'3'

4'

5'

1 2

345

6

1'2'3'

4'

5'

12

34

567

89

Pyrimidine nucleosides Purine nucleosides

Figure 1: Natural deoxynucleosides

1.1.2. Modified nucleosides in clinical use

As mimics of natural nucleic acid constituents, modified nucleosides can have a high impact on

fundamental processes involving the replication and transcription of our genetic material. This

characteristic was the basis for the development of therapies against hyperproliferative diseases –

6 Introduction

including viral infections and cancer - that are particularly dependent on high levels of nucleic acid

synthesis.

Nucleosides with antiviral activity

A number of nucleoside analogues were found to exert antiviral activity by inhibiting viral DNA

polymerases, e.g. the DNA and RNA dependent reverse transcriptase. The discovery of zidovudine

(azidothymidine, AZT) (Figure 2) as anti HIV compound in 1985 (Mitsuya et al. 1985) marked a

breakthrough in the therapy of HIV infections. Indeed AZT, later marketed as “Retrovir” by

GlaxoSmithKline, was the first drug approved by the FDA for the treatment of HIV infection. The

disclosure of other nucleoside analogues with anti-HIV activity followed (Figure 2). The common

feature is the absence of the 3′-hydroxyl group that is responsible for the chain terminator effect

within reverse transcription.

O

N3

HO

NH

N

O

O

Zidovudine Zalcitabine Didanosine Stavudine

Lamivudine Abacavir Emtricitabine

H3C

OHO

N

N

NH2

OO

HO

NH

NN

N

O

OHO

NH

N

O

O

H3C

S

O

HO

N

N

NH2

O

OHO

N

NN

N

NH

NH2S

O

HO

N

N

NH2

O

F

Figure 2: Nucleoside reverse-transcriptase inhibitors in current clinical use for the treatment of HIV infections. While zidovudine marks the oldest anti-HIV drug (launched in 1987), emtricitabine represents the last drug approved by the FDA in (2003). For further details on trade names and companies see (Flexner 2007).

Today, nucleoside reverse transcriptase inhibitors play a crucial role as components of the so-called

highly active antiretroviral therapy (HAART) that has dramatically improved the quality of life and

prognosis of patients infected by HIV. The medication includes a combination of two or more

nucleoside reverse transcriptase inhibitors and protease inhibitors (Murphy et al. 2001). Other

modified nucleoside analogues are in clinical use for treatment of infections with Hepatitis B, Herpes

Simplex, Varicella-Zoster and Hepatitis C virus (De Clercq 2004).

The viral polymerase is one of the most common targets of nucleoside reverse transcriptase

inhibitors in antiviral therapy (Berdis 2008). After the incorporation into a nascent DNA chain,

artificial nucleosides prevent the formation of the 5′ to 3′ phosphodiester linkages that are essential

for DNA chain elongations. Noteworthy nucleoside reverse transcriptase inhibitors require

Introduction 7

intracellular phosphorylation to their triphosphate forms for activity – a process that is dependent on

deoxynucleosides kinases. In case of AZT the phosphorylation was found non-selective. Conversely

the azidothymidine triphosphate competed about 100-fold better for the HIV reverse transcriptase

than for the cellular DNA polymerase (Furman et al. 1986).

A less common strategy in antiviral therapies constitutes the induction of lethal mutagenesis by the

administration of promutagenic nucleoside analogues. The rationale behind is to take advantage of

the error-prone viral DNA synthesis, while the cellular DNA polymerase is characterized by higher

fidelity and proofreading capacity. By a careful selection of the dose, the promutagenic nucleosides

would thus have only little impact on the integrity of the cellular genome, while the viability of the

virus is negatively affected. An example of such an antiviral nucleoside drug is Ribavirin that is used

for the treatment of Hepatitis C (Berdis 2008).

Despite the impressive progress in antiviral therapy, the emergence of drug-resistant mutants and

adverse side effects of many nucleoside analogues calls for the development of new drugs that may

complement the currently used ones. Examples of potential future nucleoside drugs are 2′-deoxy-4′-

C-ethynyl-2-fluoroadenosine and 2′-deoxy-4′-C-ethynyl-2-chloroadenosine. Both were reported to be

highly active against multi-drug resistant HIV and exert favourable toxicity profiles (Ohrui 2011).

Modified nucleosides as anticancer agents

Similar to viral infections, cancer is characterized by uncontrolled DNA synthesis and is therefore

conceived as hyperproliferative disease. Nucleoside reverse transcriptase inhibitors exert their

antiviral activity predominantly directly by chain termination. By contrast, nucleoside anticancer

agents typically inhibit additionally also other enzymes involved in the metabolism of nucleic acid

constituents. A prominent additional target for example is ribonucleotide reductase. DNA damage,

provoked by different factors, will eventually lead to the activation of signalling pathways that

initiate apoptotic processes. Hence apoptosis is considered as the final outcome of the treatment

with nucleoside anticancer agents (Sampath et al. 2003).

Examples of nucleoside analogues with anticancer activity currently in clinical use are shown in

Figure 3. A remarkable case study represents the discovery of Clofarabine that was approved in 2004

by the FDA for the treatment of acute lymphoblastic leukaemia in paediatrics and is marketed as

“Clolar” by Genzyme. Clofarabine can be considered as second-generation nucleoside anticancer

agent and is closely related to Cladribine and Fludarabine - clinical drugs for the treatment of

leukaemia. Both compounds are adenosine analogues and hence quite similar from the structural

point of view. Nevertheless, there are distinct advantages and disadvantages of both drugs resulting

from the nature of the specific substituents. The advantage of Cladribine lies in the 2-chloro

substitution of the purine ring, that confers improved stability against adenosine deaminase (Carson

et al. 1980).

8 Introduction

Cytarabine Gemcitabine

Cladribine Fludarabinephosphate Clofarabine

O

HO

HO

N

N

NH2

O

OH

O

HO

HO

N

N

NH2

O

F

F

Acute myeloid leukemia (Pfizer) Pancreatic and lung cancer (Eli Lilly)

O

N

NN

N

NH2

HO

HO Cl

Hairy cell leukemia (Pfizer)

O

N

NN

N

NH2

HO

OH

O FP

O

HO

OH

Chronic lymphocytic leukemia (Sandoz)

O

N

NN

N

NH2

HO

F

HO Cl

Refractory acute lymphoblastic leukemiain pediatrics (Genzyme)

Figure 3: Modified nucleosides in anticancer therapy

Even though the 2-fluoro substitution present in Fludarabine has a similar effect, the 2-chloro

substitution is preferred due to the severe toxic effect of the hydrolysis product 2-fluoroadenine that

may evolve from Fludarabine in the cell (Bonate et al. 2006). The disadvantage of both drugs is the

susceptibility to glycosyl bond cleavage. In case of Fludarabine the instability is caused by nucleoside

phosphorylase activity while in case of Cladribine both hydrolytic and enzymatic activity is

responsible. Owing to the arabino configuration of the C2′ hydroxyl group, the situation in

Fludarabine is better than in Cladribine. Nevertheless, it was later shown that the stability of the

glycosyl bond is even further improved by the presence of a fluorine atom in the C2′ arabino position,

which conferred resistance against nucleoside phosphorylase degradation (Montgomery et al. 1986).

Finally Montgomery and co-workers synthesized a number of 2′-fluoro-2-halo derivatives of 9-β-D-

arabinofuranosyladenine and found that 2-fluoro, 2-bromo and 2-chloro substituted derivatives of 9-

(2-deoxy-2-fluoro--D-arabinofuranosyl)adenine showed anti-leukaemic activity in the mouse model

(Montgomery et al. 1992). Following studies showed that the 2-chloro congener, later assigned as

Clofarabine, exerted the best activity (Bonate et al. 2006).

Despite the discovery of numerous anticancer agents, there is still the need for novel and advanced

drugs. A driving force for future developments lies in the poor selectivity of many topical

chemotherapeutic drugs that leads to the damage of healthy cells and organs. Moreover, multidrug

resistance emerges after prolonged incubation, caused for example by the activation of

transmembrane proteins effluxing active agents from the cell (Lowenthal and Eaton 1996,

Stavrovskaya 2000). A recently developed nucleoside drug with anticancer activity is Sapacitabine

that is now in clinical trial. Sapacitabine was demonstrated to be expedient for the treatment of solid

Introduction 9

tumours and haematological malignancies and was reported to have the potential of overcoming

resistance to some of the currently used drugs (Liu et al. 2012, Serova et al. 2007). Another objective

is to target nucleoside-transporter deficient cells that are highly resistant to nucleoside analogues.

Such challenges might be tackled by encapsulation or conjugation of target compounds to

nanoparticles (Hajdo et al. 2010). The development of suicide gene therapies represents an attempt

towards the selective killing of tumour cells. Parker and co-workers have demonstrated the feasibility

of a system employing modified nucleosides as prodrugs that are selectively cleaved in E. coli purine

nucleoside phosphorylase (PNP) transfected cancer cells. With this method cytotoxic purine bases as

for example 2-fluoroadenine could be locally liberated (Parker et al. 1997, Parker et al. 2003).

Other fields of pharmaceutical application

The potential of modified nucleoside drugs is not restricted to the treatment of viral infections and

cancer. In fact, modified nucleosides that act as purine nucleoside inhibitors are known for their

therapeutic properties. A recent example in this field constitutes an immucillin purine nucleoside

phosphorylase inhibitor that proved to be potent in a primate animal model for therapy against the

protozoan parasite Plasmodium falciparum. The parasite is responsible for most of the malarial

deaths each year (Cassera et al. 2011).

Moreover, insights into the biological activities on adenosine receptors have further broadened the

scope of potential applications of adenosine analogues. A number of adenosine analogues acting as

adenosine receptor agonists are now in clinical trials and may later be used for the treatment of

inflammation, type 2 diabetes, and arrhythmia (Samsel and Dzierzbicka 2011).

1.1.3. Modified nucleosides as building blocks for synthetic nucleic acids

The modulation of gene expression through the use of synthetic nucleic acids is an exciting research

field in both fundamental and clinical science. Antisense oligonucleotides (ASOs) and small

interfering RNAs (siRNAs) are the most widely used strategies. The profound impact on clinical drug

development is reflected by the numerous oligonucleotides that are currently in clinical trials (Watts

and Corey 2010, Watts and Corey 2012). Fomivirsen is the first ASO approved by the FDA and is used

for the treatment of cytomegalovirus retinitis (Grillone and Lanz 2001).

Pioneering studies on antisense ASOs have been reported in 1978 by Paul Zamecnik (the “Father of

Antisense”(Agrawal 2010)) and Mary Stephenson. The idea was to administer an artificial

oligonucleotide complementary to a target RNA into cells and thereby inhibiting the expression of a

target gene. The work published by Zamecnik and Stephenson demonstrated the feasibility of the

approach for the inhibition of Rous sarcoma virus replication in fibroblast cultures (Stephenson and

Zamecnik 1978, Zamecnik and Stephenson 1978).

A more recent development is the use of siRNAs that are double stranded RNA oligonucleotides

capable of entering the RNA interference pathway naturally occurring in the cell. The manipulation of

gene expression in the nematode Caenorhabditis elegans by taking advantage of RNA interference

was reported in 1998 (Fire et al. 1998). In 2006 Andrew Fire and Craig C. Mello were rewarded for

10 Introduction

their groundbreaking contribution with the Nobel Prize in Physiology or Medicine. RNA interference

(RNAi) is characterised by the cleavage of double stranded RNAs into small RNAs that afterwards

associates with proteins to form a RNA induced silencing complex. In the subsequent process the

sense strand is released, while the antisense siRNA is used to identify and destroy the homologous

mRNA target (Hammond et al. 2000, Martinez et al. 2002).

Modified nucleosides in antisense oligonucleotides and siRNA

In both antisense and RNA interference technologies, early optimism was significantly diminished

after a multitude of severe hurdles were discovered (Gura 1995). In particular, the following

obstacles and requirements have been identified as bottlenecks: i) instability of the oligonucleotides

due to nuclease-mediated degradation, ii) poor selective and stable binding to the target RNA, iii) un-

wanted triggered immune responses and iv) limitations on the level of delivery and binding to other

proteins.

Soon it was found that chemical modifications of the nucleotides could convey favourable properties

to tackle some of the aforementioned limitations. Thus, in first-generation oligonucleotides,

phosphodiester linkages were replaced by phosphorothioate linkages in order to prevent the

degradation by nucleases. Among other modified building blocks, 2′-fluorinated nucleosides proved

to have unique, favourable properties for antisense and RNA interference.

O

HO F

HO

NH

N

O

O

O

HO

F

HO

NH

N

O

O

Figure 4: Uridine substituted with fluorine in the 2′-ribo position (left) and 2′-arabino position (right)

The substitution of the 2′-ribo position with a fluorine atom locks the sugar moiety predominantly in

a C3′-endo conformation that is characteristic for the sugars in RNA helices. This feature was

exploited for the development of ASOs that better bind to the RNA target molecule due to increased

thermodynamic stability of the generated DNA/RNA duplex (Kawasaki et al. 1993). Likewise many

studies have demonstrated the favourable properties of 2′-fluorinated nucleotides in RNA

interference (Allerson et al. 2005, Deleavey et al. 2010, Manoharan et al. 2011, Morrissey et al.

2005). Particularly it was reported that the 2′-fluoro substitution leads, in comparison to non-

modified oligonucleotides, to constructs with enhanced serum and thermal stability of the duplex

with reduced immunogenicity. Furthermore, higher in vitro and in vivo potency, also in comparison

with some other modified oligonucleotides (including locked nucleic acids) was reported (Manoharan

et al. 2011).

Introduction 11

Replacement of the 2′-arabino position of the sugar with a fluorine atom leads to equally favourable

properties for RNA technology. The recruitment of RNase H that cleaves RNA in RNA/DNA duplexes

improves the efficacy of many ASOs (Watts and Corey 2012). However, chemical modifications that

are conventionally used to increase the stability of the RNA/DNA duplex or confer nuclease

resistance do not support RNase activity. In order to achieve nevertheless RNase susceptibility,

typically “gapmers” are used in which unmodified nucleotides are introduced in the middle of

oligonucleotides containing also modified constituents (Monia et al. 1993, Watts and Corey 2012).

Also, it was shown that 2′-fluoroarabinonucleic acid constituents supported RNase H activity while

retaining a high binding affinity to the RNA target (Damha et al. 1998, Watts and Damha 2008).

Hence 2′-fluoroarabinonucleosides are attractive components of oligonucleotides used for gene

silencing (Kalota et al. 2006). Recently, the expedient properties of oligonucleotides with N3′-P5′

phosphoramidate linkages (Gryaznov et al. 1995, Gryaznov 1999) were combined with the

advantages of the fluorination of the 2′-arabino position of the sugar residue. The synthesis and use

of the resulting 2′-arabino-fluorooligonucleotide N3′-P5′ phosphoramidates was patented (Gryaznov

and Schultz 2011).

Modified nucleosides for the stabilization of aptamers

The binding of oligonucleotides to other proteins in addition to the intended hybridization with the

target RNA is generally considered as unwanted off-target effect in antisense and RNA interference

strategies. However, the high affinity binding of single-stranded nucleic acids, to distinct molecular

targets has been found to be of interest also for other therapeutic approaches (Thiel et al. 2009). The

specificity of these “aptamers” is conferred by their three-dimensional structure. The incorporation

of 2′-fluoro pyrimidine nucleotides has been shown to increase the resistance against nucleases and

leads to equal or higher binding affinities to the target ligand (Adler et al. 2008, Khati et al. 2003). In

2004 the first aptamer therapeutic (Pegaptanib) has gained FDA approval and is now marketed by

Pfizer for the treatment of exudative (wet) age-related macular degeneration. The oligonucleotide

that is substituted with 2′-fluorinated pyrimidine nucleotides, is selectively directed against a

vascular endothelial growth factor (Gragoudas et al. 2004, Ng and Adamis 2006).

Noteworthy, current applications of 2′-fluorinated nucleosides often concentrate on 2′-fluro

substituted pyrimidine nucleosides - a phenomenon that might reflect the fact that the chemical

synthesis of the according purine nucleosides is significantly more challenging and therefore related

to higher costs. This aspect will be further discussed throughout the present study.

1.2. Synthetic routes towards nucleoside analogues

1.2.1. Chemical synthesis

For the chemical synthesis of nucleosides, typically a convergent approach is followed, which means

that heterocyclic base and ribose moiety are independently prepared and afterwards coupled. A

commonly used method is the silyl-Hilbert-Johnson (or Vorbrüggen) reaction, in which a silylated

12 Introduction

(nucleophilic) heterocyclic base reacts with a protected (electrophilic) sugar acetate in the presence

of a Lewis acid (Vorbrüggen and Ruh-Pohlenz 2001).

Even though the Vorbrüggen reaction and similar methods have been widely applied for the

synthesis of nucleosides, the chemical synthesis is often challenged by regio- and stereospecific

requirements that are prerequisites for the biological activity of nucleosides. In natural nucleosides

the carbohydrate moiety is connected to the N9 atom of purine bases and to the N1 atom of

pyrimidine bases, respectively. However, the presence of multiple nucleophilic sites of the

heterocyclic base poses difficulties for the regioselective formation of the glycosyl bond during

chemical synthesis. Moreover, natural nucleosides can be exclusively found in the β-anomeric

configuration in which the base is oriented above the plane of the sugar. Indeed, this requirement

can be readily achieved in the synthesis of ribonucleosides through the formation of a cyclic cation

intermediate of the ribose moiety prior to the formation of the glycosyl bond. For the synthesis of

deoxyribonucleosides the situation is more complicate due to the absence of the 2′ hydroxyl group

involved in the neighbouring group participation described above. More details on the specific

challenges encountered in the synthesis of 2′-fluorinated and 2,6-dihalogenated nucleosides will be

discussed in section 5.1.

In summary, the chemical synthesis is typically a multistep process requiring protection and

deprotection of functional groups and sophisticated procedures to achieve regio- and

stereoselectivity. Furthermore, the chemical preparation usually involves reagents that are harmful

to health and environment.

1.2.2. Enzymes for nucleoside synthesis

Enzymes are able to efficiently catalyze reactions with strict regio- and stereoselectivity under mild

reaction conditions. Protection and deprotection steps, as well as the employment of hazardous

chemical reagents are unnecessary. If utilized as biocatalysts, enzymes have thus the potential to

replace complicate chemical reactions routes. A number of enzymes interacting with nucleic acid

constituents have been investigated with respect to their ability to aid in nucleoside synthesis. It was

found that enzyme catalyzed reactions can be exploited in two major fields: i) the selective

modification of nucleosides, and ii) the formation of the glycosyl bond connecting heterocyclic base

and pentofuranose moiety.

The hydrolytic deamination of 6-aminopurine, catalysed by adenosine deaminase, represents an

example for the first category of enzymes and has been widely applied for the synthesis of modified

purine nucleosides (Santaniello et al. 2005). Moreover, enzymes have been used for the selective

modification of the sugar moiety of nucleosides. Examples include lipases, used for the regioselective

acylation of hydroxyl groups (Moris and Gotor 1993), and nucleoside oxidases that oxidize the

CH2OH group of the sugar moiety of nucleosides, which can be used for the synthesis of carboxylic

nucleoside derivatives (Mahmoudian et al. 1998).

Introduction 13

The second major application field of enzymes in nucleoside chemistry concerns the formation of the

glycosyl bond between the heterocyclic base and the pentofuranose moiety. In practice this strategy

involves the transfer of a glycosyl residue from a nucleoside donor to an acceptor base and is

catalysed by NPs or N-deoxyribosyltransferases (NdRT). Although the overall reaction – the

interchange of the base of a nucleoside - is essentially the same (Figure 5), the catalytic mechanism

of the transglycosylation reaction mediated by NPs and NdRTs is different. NdRTs (EC 2.4.2.6)

catalyze the direct transfer of a deoxyribofuranosyl moiety, while with NPs the intermediate product

α-D-pentofuranosyl-1-phosphate is formed. For this reason the NP catalyzed reaction requires the

presence of inorganic phosphate. Furthermore, NPs and NdRTs differ in their substrate specificities.

Both ribo- and deoxyribonucleosides are natural substrates of nucleoside phosphorylases. Contrarily

NdRTs are specific for 2′-deoxyribonucleosides. Regarding the specificity toward the heterocyclic

bases, NdRTs are classified in two categories. Type I is specific for the interchange of purine bases

(purine <-> purine) whereas type II NdRTs catalyze the transfer of purine and pyrimidine bases

(purine <-> purine, pyrimidine <-> pyrimidine, purine <-> pyrimidine) (Holguin and Cardinaud 1975).

In particular NdRTs from Lactobacilli have been extensively studied and used for the synthesis of

natural and artificial nucleosides (Carson and Wasson 1988, Fernandez-Lucas et al. 2010, Huang et al.

1983, Kaminski 2002, Okuyama et al. 2003).

Remarkably, metabolic enzymes of purine and pyrimidine nucleotide metabolism have also been

used to completely reshape the synthesis of nucleosides in vitro. Even such an approach appears

tedious for production of bulk chemicals, it is an important tool for the synthesis of isotope labelled

nucleic acid constituent (Schultheisz et al. 2008, Schultheisz et al. 2010).

O

HO OH

HO

NH

N

O

O

Pyrimidine nucleoside(Pentofuranosyl donor)

Pyrimidine basePurine base (pentofuranosyl acceptor)

Purine nucleoside

N

NNH

N

NH2

NH

NH

O

O

O

HO OH

HO

N

NN

N

NH2

Figure 5: General scheme of a transglycosylation reaction with a pyrimidine nucleoside as pentofuranosyl donor and a purine base as pentofuranosyl acceptor. The overall reaction can be catalyzed either by a single type II N-deoxyribosyltransferase or by two NPs in the presence of inorganic phosphate. As example the synthesis of adenosine from uridine as pentofuranosyl donor and adenine as pentofuranosyl acceptor is shown.

14 Introduction

1.2.3. The NP-catalyzed transglycosylation of nucleosides

The transglycosylation of nucleosides, mediated by nucleoside phosphorylases, proceeds in two

consecutive steps (Figure 6). In the first step a nucleoside that serves as pentofuranosyl donor is

phosphorolytically cleaved into the corresponding heterocyclic base and α-D-pentofuranosyl-1-

phosphate. In a second step this activated carbohydrate moiety is coupled to the heterocyclic base

that is used as pentofuranosyl acceptor. Hence, in the first reaction inorganic phosphate is

consumed, while in the second reaction phosphate is released.

Both reactions are catalyzed by NPs. Depending on the type of NPs used different modes of

transglycosylations are possible. If a purine nucleoside phosphorolyzing enzyme is exclusively

employed as biocatalyst, interchange of purine bases can be achieved. For this purpose

7-methylguanosine and 7-methylinosine have been reported as effective ribofuranosyl donors

(Hennen and Wong 1989, Ubiali et al. 2012). Likewise the interchange of pyrimidine bases can be

accomplished by utilizing a pyrimidine phosphorolyzing enzyme in a transglycosylation with a

pyrimidine nucleoside serving as pentofuranosyl donor and a different pyrimidine base serving as

pentofuranosyl acceptor. Following this methodology Serra and co-workers have synthesized 5-

fluoro-2′-deoxythymidine with immobilized E. coli thymidine phosphorylase (Serra et al. 2011).

Pi

+

Uridine

Phosphate

Pi

Uracil α-D-pentofuranose-1-phosphate

Adenine

Phosphate

PyNP

PyNP : Pyrimidine nucleoside phosphorylase

PNP : Purine nucleoside phosphorylase

O

HO OH

HO

NH

N

O

O NH

NH

O

O

O

HO OH

HO

OPO32-

N

NNH

N

NH2

O

HO OH

HO

N

NN

N

NH2

PNP

Figure 6: Purine nucleoside synthesis at the example of uridine as pentofuranosyl donor and adenine, functioning as pentofuranosyl acceptor

The spectrum of possible transglycosylation reactions is expanded through the combined use of

pyrimidine and purine nucleoside phosphorolyzing enzymes. With this strategy a pyrimidine base can

be substituted by a purine base or vice versa. An example is the synthesis of 5′-methyluridine by

employing guanosine as pentofuranosyl donor and thymine as acceptor (Gordon et al. 2011, Ishii et

al. 1989). The advantage is that guanine, the phosphorolysis product of guanosine precipitates from

the reaction and thereby shifts the equilibrium of the first reaction in the favourable direction.

However, of particular interest for the present study is the synthesis of modified purine nucleosides

Introduction 15

accomplished by a transglycosylation reaction with pyrimidine nucleosides acting as pentofuranosyl

donor and purine bases acting as pentofuranosyl acceptor (Figure 6). The rationale behind this

approach is that a number of modifications on the carbohydrate moiety of pyrimidine nucleosides

can be relatively easily introduced by chemical transformations. The procedure involves the

intermediate formation of an intramolecular anhydrous bond between the O-2 atom of the

pyrimidine base and the 2′, 3′, or 5′ position of the carbohydrate moiety and subsequent opening of

the oxygen containing bridge by the treatment with nucleophilic agents (illustrated in (Mikhailopulo

and Miroshnikov 2011)). By contrast, comparably simple methods for the modification of the

carbohydrate moiety of purine nucleosides do not exist. The coupling of sugar modified moieties,

donated by pyrimidine nucleosides, to natural or chemically prepared purine bases is hence an

attractive approach for the synthesis of purine nucleoside analogues (Krenitsky et al. 1981,

Lewkowicz et al. 2000, Mikhailopulo 2007, Tuttle et al. 1993, Utagawa et al. 1985b, Utagawa 1999).

1.2.4. Formulation of NPs for technical use

The least laborious way to exploit the enzymatic activities of NPs in enzymatic transglycosylations of

nucleosides is the utilization of whole cells displaying the desired activities. With this strategy

downstream processing is virtually not required which leads to a cost-effective biocatalyst

preparation. Moreover, the whole cell represents a kind of natural immobilization vehicle which

simplifies the handling and conveys stability to the biocatalysts. On the downside, whole cell

biocatalysts are complex systems exhibiting not only NP activities but various other catalytic

reactions. The overall process can hence become complicate. An example represents the synthesis of

adenosine from uridine and adenine catalyzed by E. coli BL21, as described by Lewkowicz and co-

workers (Lewkowicz et al. 2000). Even though the authors reported a 94 % yield of adenosine after

only one hour, they also observed that after prolonged incubation times the second (PNP catalyzed)

reaction was reversed and after 24 h only adenine and uracil were present. A possible explanation is

that α-D-pentofuranose-1-phosphate served as energy source for the cells and has been consumed.

Adenosine deaminase activity is another by-activity of whole cells that is often encountered as

obstacle. Comparative investigations of E. coli “adenosine nucleoside phosphorylase” (today referred

to as EcPNP) and adenosine deaminase, have shown that the latter is more sensitive to heat (Koch

and Vallee 1958). Investigations of Utagawa and co-workers revealed later that conducting the

enzymatic synthesis of ara-A at high temperature (60 °C instead of 37 °C) was a good strategy to

avoid adenosine deaminase activity of the Enterobacter aerogenes cells employed as biocatalyst

(Utagawa et al. 1980).

In other cases, by-activities of whole cells are rationally used to pursue the desired synthetic

direction. For example, cytidine deaminase activity of selected E. coli cells was used to transform 1-β-

D-arabinofuranosyl cytidine and 2′-deoxy-2′-fluorocytidine into the respective uridine derivatives,

prior to the transglycosylation reactions affording 1-β-D-arabinofuranosyl guanine and 2′-deoxy-2′-

fluoroguanosine (Mikhailopulo 2007, Zaitseva et al. 1999).

16 Introduction

The use of purified enzyme preparations offers the possibility to apply very high enzyme loadings

which is in some cases required due to poor substrate activities of chemically modified precursors

(Tuttle and Krenitsky 1992, Tuttle et al. 1993). Furthermore, purified enzymes can be selectively

applied and side-reactions as described above are not of concern.

In view of industrial applications however, it is highly desirable to make use of immobilized enzyme

preparations instead of directly employing purified enzyme solutions. Immobilized biocatalysts can

be easily recovered from the reaction mixture which simplifies the downstream processing and

facilitates biocatalyst recycling. Furthermore, immobilization is a useful tool to increase the stability

of enzymes to withstand harsh reaction conditions. In this regard the preservation of the multimeric

nature of NPs deserves special attention.

Diverse methods of immobilization have, therefore, been exploited for the use of nucleoside

phosphorylases as biocatalysts. Earlier studies report on the successful co-immobilization of NPs on

anion exchange resins (DEAE) (Hori et al. 1991, Mahmoudian 2000). To further increase chemical

stability, Zuffi and co-workers have co-immobilized E. coli UP and PNP by covalent linkage on epoxy

activated Sepabead®s resins (Zuffi et al. 2004). The higher stability allowed the use of the biocatalyst

at relatively high temperature (60 °C) in the presence of organic solvents (40 % DMSO). Rocchietti

and co-workers found the covalent linkage on glyoxyl-agarose as most suitable for B. subtilis PNP

immobilization (Rocchietti et al. 2004). Likewise PNP and PyNP from G. stearothermophilus have

been successfully covalently immobilized on aminopropylated macroporous glass (Taran et al. 2009).

On the other hand, there seems to be evidence that ionic adsorption is advantageous over covalent

immobilization for certain NPs. This appears to apply especially for members of the NPII family,

which are characterized by a homodimeric structure and a domain movement within catalysis

(Pugmire and Ealick 2002). Thus for both the immobilization of B. subtilis PyNP and E. coli TP, ionic

adsorption was found to be a suitable method while covalent immobilization was generally

detrimental (Rocchietti et al. 2004, Serra et al. 2011). A completely different strategy was followed

by Visser and co-workers. They applied spherezyme selfimmobilization (Brady et al. 2008) to further

increase the stability of an E. coli UP derivative that gained improved thermal stability through

directed evolution (Visser et al. 2011).

1.3. Nucleoside phosphorylases

The reversible phosphorolysis of nucleosides by nucleoside phosphorylases was first described by

Kalckar in 1947. Since then these remarkable enzymes have been in focus of numerous research

endeavours. The following sections will give a brief overview of some of the lessons learned.

1.3.1. Physiological role

Nucleoside phosphorylases catalyze the reversible phosphorolysis of ribo- and deoxyribonucleosides

in the presence of inorganic phosphate. The products of the reaction are α-D-pentofuranose-1-

phosphate and the nucleobase of the nucleoside substrate. In vitro studies revealed that the

Introduction 17

equilibrium of the phosphorolysis reaction is shifted towards the reverse (synthetic) reaction,

whereby this effect is significantly more pronounced in PNPs than in pyrimidine nucleoside

phosphorolyzing enzymes. Thus the equilibrium constant of the phosphorolysis reaction determined

for E. coli uridine phosphorylase with uridine as substrate was in the range of 0.54 – 0.61 (Vita et al.

1983), while the equilibrium constant obtained with E. coli PNP with inosine as substrate was 0.0175

(Jensen and Nygaard 1975). Due to these findings, PNPs have long been considered as enzymes

involved in the salvage of purine bases. However, in vivo the phosphorolysis reaction is highly

preferred over the synthetic reaction, since the products are further metabolized. Purine bases are

salvaged by hypoxanthine-guanine phosphoribosyltransferease or oxidized by xanthine oxidase to

uric acid (Bzowska et al. 2000), whereas the liberated pentose-1-phosphate is used in the catabolism

as energy source (Sgarrella et al. 1997). The catabolic role of nucleoside phosphorylases is also

reflected by the fact that they often belong to the same regulon as other nucleoside catabolising

enzymes (Hammer-Jespersen and Munch-Ptersen 1975, Tozzi et al. 1981).

However, in some specific cases NPs are used for the salvage of nucleobases, and thus permit an

alternative to the de novo synthesis of nucleosides. For example many protozoan parasites lack the

de novo purine nucleoside synthesis and rely on NPs instead to obtain building blocks for nucleic acid

synthesis (Hammond and Gutteridge 1984). This phenomenon makes the PNP of the malaria parasite

Plasmodium falciparum a promising target for the development of anti malaria drugs (Silva et al.

2007).

In humans, PNP deficiency was found to coincide with defected T-cell immunity, while B-cell function

was unaffected (Giblett et al. 1975). This discovery was the driving force for extensive efforts

towards the development of PNP inhibitors that were now recognized as potential

immunosuppressive agents for organ transplantation, as well as for the treatment of T-cell leukaemia

and T-cell related autoimmune diseases (Bzowska et al. 2000, Silva et al. 2007). Noteworthy, human

thymidine phosphorylase was found to be an angiogenic factor and proved to be identical to the

platelet-derived endothelial cell growth factor (Akiyama et al. 2004, Furukawa et al. 1992).

1.3.2. Basic nature of the catalytic mechanism

The basic principle of the catalytic mechanism is presumably similar in all nucleoside phosphorylases

(Pugmire and Ealick 2002). Studies on human PNP revealed that the nucleoside is bound in a high

energy conformation, producing steric strain that is favourable for the glycosidic bond cleavage.

Through electron flow from O4′ of the pentose moiety to the purine ring, an oxocarbenium ion is

formed, stabilized by the negative charge of the phosphate ion. Finally, the phosphate ion

participates in a nucleophilic attack at the C1 position. Active site residues interactions at the N-7

position are likely to support the flow of electrons from the glycosidic bond to the purine ring.

Active sites in E. coli and human PNP

As it will be discussed in more detail in the following section (1.3.3) nucleoside phosphorylases of the

NP-I fold encompass bacterial and eukaryotic type PNPs that are characterized by distinct substrate

18 Introduction

specificities. In both type of enzymes a ternary complex is formed which involves the enzyme and

both substrates (Bzowska et al. 2000). Moreover the geometric arrangement of purine, ribose and

phosphate binding site is similar (Mao et al. 1997). However, the mode of substrate binding differs

significantly and may account for the different substrate specificities of eukaryotic PNPs (specific for

6-oxopurines) and bacterial PNPs (accepting both 6-oxopurines and 6-aminopurines). The following

paragraphs are mostly based on (Bzowska et al. 2000) and will shortly list some differences with the

example of EcPNP and human PNP.

Figure 7: Proposed catalytic mechanism of hPNP. Figure adopted from (Pugmire and Ealick 2002)

A basic difference of the active site of both type of enzymes, is that in hexameric (bacterial type) PNP

the active site is composed out of amino acids belonging to two subunits: In EcPNP His4 and Arg43

are donated by the neighbouring subunit and interact directly via hydrogen bonds with the ligands

(Figure 8). By contrast, in hPNP Phe159 is the only residue from the neighbouring subunit and does

not directly interact with ligands. Its role is rather to complete the hydrophobic environment near

the pentose group.

In EcPNP the base binding site is more exposed and accessible than in hPNP. The N1-H of the purine

base is involved in a hydrogen bond with Glu201 in hPNP, whereas in EcPNP N1-H is linked to a water

molecule and has no contact to a protein residue. In 6-Aminopurines the aminogroup interacts with

Asp204 in EcPNP, which plays obviously a key role, since it is conserved in enzymes homologous to

EcPNP. In hPNP Glu201 and Asn243 belong to the essential residues involved in base binding. The

role of Asn243 may be attributed to the recognition of 6-oxo purines as substrates: The replacement

of Asn243 by Asp changed specificity of trimeric PNPs from 6-oxopurines to that of bacterial type

PNPs.

The phosphate binding site is more positively charged in EcPNP than hPNP, which is the result of

three Arg residues interacting with phosphate (Arg24, Arg87, Arg43), whereas in human PNP there is

only one Arg residue (Arg84) that interacts with phosphate.

Both PNP types use hydrophobic interactions for the hydrophobic site of pentose ring, while

hydrophilic site of the pentose ring, with hydroxyl groups, faces the phosphate binding site.

Introduction 19

Figure 8: Key active site residues of EcPNP and their interactions with inosine. W = water molecule. Figure taken from (Bennett et al. 2003)

Overview of different nucleoside phosphorylases

All nucleoside phosphorylases are pentosyltransferases that catalyze the β-glycosidic bond cleavage

of nucleosides, whereby a pentufranosyl-1-phosphate is formed under inversion of the configuration.

Based on the specific substrates, enzyme names and enzyme commission number (EC) were

assigned. Table 1 lists enzymes and acronyms that will be of relevance in the present report.

Table 1. Nucleoside phosphorylases

Enzyme name Acronym EC number

Purine nucleoside phosphorylase PNP 2.4.2.1

Pyrimidine nucleoside phosphorylase PyNP 2.4.2.2

Uridine phosphorylase UP 2.4.2.3

Thymidine phosphorylase TP 2.4.2.4

5′ -Methylthioadenosine phosphorylase MTAP 2.4.2.28

20 Introduction

1.3.3. Classification of NPs

In 2002 a comprehensive classification of nucleoside phosphorylases was established by Pugmire and

Ealick. It was found that NPs can be assigned to one of two groups that are distinct in the structural

fold of their subunits (Figure 9). The nucleoside phosphorylase-I family is characterized by a common

single domain α/β subunit fold with trimeric or hexameric quaternary structure and includes NPs

accepting purine nucleosides (viz. PNP and MTAP) as well as uridine phosphorylase (UP). Members of

the NP-II family are homodimers, in which each subunit consists out of two domains: a large mixed

α/β domain, separated by a cleft from a smaller α-helical domain. NPs assigned to this family share a

high degree of sequence identity and are currently known to exclusively accept pyrimidine bases and

corresponding nucleosides as substrate but not purine bases or purine nucleosides.

NP-I family

• Subunits with common α/β-fold (single-domain)

• Substrates: Purine nucleosides and uridine

Hexamers Trimers

NP-II family

• Homodimeres, two-domain subunits

• High degree of sequence identity

• Significant domain movement needed for catalysis

• PyNP: phoyphorolyses of thymidine and uridine

• TP: specific for thymidineBacterial PNP UP

Specific for:

6-aminopurines

6-oxopurine

Mammalian PNP MTAP

Specific for:

6-oxopurines

Figure 9: Classification of NPs based on Pugmire and Ealick (Pugmire and Ealick 2002)

NP-I family

On the basis of substrate specificity, molecular mass, quaternary structures and amino acid

sequences, members of the NP-I family can be further categorized. Hence Pugmire and Ealick have

defined a trimeric and a hexameric subgroup of the NP-I family. PNPs with trimeric quaternary

structure, described before by Bzowska as low-molecular-mass PNP (Bzowska et al. 2000), are

specific for 6-oxopurines and their nucleosides, but do not accept 6-aminopurines or nucleosides

thereof (e.g. adenosine) as substrate. By contrast, high-molecular-mass PNPs (with hexameric

quaternary structure) accept both 6-aminopurines as well as 6-oxopurines as substrate. However, in

some cases high-molecular-mass PNPs were found to have a significant higher specificity towards

adenosine in comparison to 6-oxopurine (nucleosides) (Mcelwain et al. 1988, Sgarrella et al. 2007,

Trembacz and Jezewska 1993). Therefore, respective enzymes are also referred to as adenosine

Introduction 21

phosphorylases, as it is the case for the high-molecular-mass PNP from B. cereus (Sgarrella et al.

2007).

While trimeric PNPs are found in eukaryotes, the hexameric form is prevalent in bacteria. However,

in a number of bacteria both, a PNP with substrate specificity of low-molecular-mass as well as a PNP

with substrate specificity for high-molecular-mass can be found. Examples include the PNPs from

E. coli, G. stearothermophilus, B. subtilis and B. cereus (Bzowska et al. 2000).

Sequence analysis has shown that the trimeric subgroup of the NP-I family also encompasses

5′-methylthioadenosine phosphorylase (2.4.2.28) and the hexameric NP-I subgroup includes uridine

phosphorylase (2.4.2.3).

NP-II family

TP and PyNP form the nucleoside phosphorylase-II family, which share a common two-domain

subunit fold and a high level of sequence identity (Pugmire and Ealick 2002). Despite the similarity of

the reaction catalyzed, uridine phosphorylase (UP; EC 2.4.2.3) belongs to the phosphorylase-I family

with distinct structural characteristics. From the catalytic point of view, TP is distinguished from UP

due to its high specificity for the 2′-deoxy-D-ribofuranose moiety of pyrimidine nucleosides (Pugmire

and Ealick 2002). By contrast, PyNP does not discriminate between uridine and thymidine and

accepts both compounds as natural substrates (Saunders et al. 1969).

Other relevant aspects concerning the classification of NPs

Ten years later, the classification of nucleoside phosphorylases presented by Pugmire and Ealick in

2002 is still helpful for the classification of NPs or for assumptions regarding quaternary structures or

substrate specificities of unknown NP gene products. However, not all relevant aspects have been

covered by the theory. Some critical aspects will be shortly stated in this section.

A number of PNPs with dimeric or tetrameric quaternary structures were reported (reviewed in

Bzowska et al. 2000), while in the classification in 2002 exclusively trimeric and hexameric quaternary

PNP structures are considered. However, many of the earlier studies have relied on gel filtration or

electrophoresis in order to determine quaternary structures. Such methods may lead to incorrect

conclusions, as illustrated by Bzowska and co-workers (Bzowska et al. 2000): By gel filtration they

found that PNP from Cellulomonas has a tetrameric quaternary structure, even though a trimeric

structure was expected (Tebbe et al. 1997). However, in the crystal structure the enzyme was later

found to be a trimer (Tebbe et al. 1999), as it was originally expected from the PNP subgroup

classification.

In other cases it appears to be obvious that quaternary structures are not consistent with the

theoretically expected structures. For example human UP was found to be a homodimer (Roosild et

al. 2009), even though the classification implies that UPs are hexamers. Likewise the crystal structure

of trypasonomal UP revealed a homodimeric enzyme (Larson et al. 2010). Moreover, a number of

hexameric PNPs were found to display substrate specificities of the low-molecular-mass PNPs that

were previously associated by Pugmire and Ealick with a trimeric quaternary structure. Examples of

22 Introduction

these hexameric enzymes are the PNPs from T. thermophilus (Tahirov et al. 2004),

Plasmodium falciparum (Daddona et al. 1986, Schnick et al. 2005, Shi et al. 2004), and from the

hyperthermophilic archaeon P. furiosus (Cacciapuoti et al. 2007).

Other deviations from the general perspective on NPs concern the substrate spectra. For example

the NP from Klebsiella was shown to accept both pyrimidine nucleosides as well as purine

nucleosides as substrates. The relative activity was highest for uridine (368 %), but the enzyme also

had high activity towards deoxyinosine (254 %) and other purine nucleosides, while thymidine

showed significant less substrate activity (29 %) (Ling et al. 1990). Other studies revealed that purine

nucleoside phosphorylases with cytidine activity exist (Mikhailopulo and Miroshnikov 2011). And

finally, a novel NP enzyme specificity was recently discovered. The putative MTAP from

Pseudomonas aeruginosa was found to be a NP with specificity towards 5′-deoxy-5′-methylinosine,

whereas 5′-deoxy-5′-methyladenosine was not accepted as substrate (Guan et al. 2011). 5′-Deoxy-5′-

methylinosine specific enzymes have not been described before.

New insights have been also gathered by comprehensive studies on arachaeal PNPs and MTAPs.

Examples include the MTAPI (Appleby et al. 2001, Cacciapuoti et al. 1994) and MTAPII from

Sulfolobus solfataricus (Cacciapuoti et al. 2005), as well as PNP (Cacciapuoti et al. 2007) and MTAP

(Cacciapuoti et al. 2004) from P. furiosus. The research in this area also opens the way for new

conclusions regarding the structural and functional differences between MTAP and PNP (Cacciapuoti

et al. 2011).

1.3.4. From natural to recombinant production of nucleoside phosphorylases

The phosphorolytic cleavage of nucleosides by NPs has been described for a multitude of living

organisms, spanning the three domains of life (Pugmire and Ealick 2002). Microorganisms have been

exploited in particular as natural nucleoside phosphorylase (NP) producers for biocatalytic

applications. In this regard, comprehensive screening studies have led to the identification of

microorganisms that are especially suitable for the enzymatic synthesis of nucleosides and analogues

thereof. Thus, Utagawa and co-workers have tested more than 240 microorganisms representing 26

bacterial genera towards their ability to produce ara-A (9-β-D-arabinofuranosyladenine) from ara-U

(9-β-D-arabinofuranosyluracil) and adenine, and selected an Enterobacter aerogenes strain as best

producer (Utagawa et al. 1980). Since the reaction did not proceed without inorganic phosphate, it

was assumed that the transglycosylation reaction was catalyzed by nucleoside phosphorylases.

B. stearothermophilus, possessing one PyNP and two PNPs was found to be the best producer of a

number of 6-modified nucleosides among 100 microorganisms screened (Trelles et al. 2005).

Likewise E. coli strains with specific catalytic properties including nucleoside phosphorylase activity

have been selected and successfully applied for the synthesis of modified nucleosides (Barai et al.

2002, Zaitseva et al. 1999, Zinchenko et al. 1990). Recently T. thermophilus strains were screened as

whole cell biocatalysts for their productivity in the synthesis of natural purine nucleosides

(Almendros et al. 2009). The T. thermophilus strains were found to have beneficial properties for this

application due to high productivity values, the absence of adenosine deaminase activity and the

Introduction 23

thermostability, which permitted to conduct the screening at 65 °C. However, based on the results it

was not clear whether NdRTs or NPs are responsible for the transglycosylation reactions.

It is obvious that such screening approaches offer a great potential for the identification of

biocatalysts with desirable substrate specificities or other favourable properties as enhanced thermal

stability. However, with the aim in view to design efficient industrial processes, the volumetric yield

of NPs reached with such strategies is often rather unsatisfactory. The reasons can be found in

difficulties to cultivate NP producer strains to high cell densities or in the low amount of NPs

naturally present in the cells. In fact, even screening approaches are often restricted to

microorganisms that are easy to cultivate in standard culture media (Condezo et al. 2006). On the

other hand, strategies to increase the expression level of NPs per cell have been developed. Thus, the

addition of nucleosides and related compounds to the culture medium was used in order to

maximize the induction of NPs in Enterobacter aerogenes (Wei et al. 2008). The extent of culture

growth, that is the growth phase of the cells at the time of harvest, constitutes another influential

parameter and was investigated to maximize the specific activity of E. coli BL21 whole cell

biocatalysts (Rogert et al. 2002).

On the other hand, the advances in recombinant DNA technology opened new dimensions of the

high-level production of proteins: By changing the genetic context of a gene coding sequence,

unnaturally high expression levels can be reached. If thereby the source species of the gene

sequence coincides with the host used to produce the target protein, the methodology is referred to

as homologous expression. With this strategy overproducing strains of E. coli UP, TP, and PNP have

been generated (Esipov et al. 2002). In contrast, heterologous expression is referred to the

methodology where a gene is derived from a species that does not coincide with the expression host,

into which the gene is inserted for overexpression. The availability of such a technology has greatly

expanded the scope of NPs that can be efficiently produced and studied. This includes NPs from

microorganisms that are difficult to cultivate, currently not available, or even not cultivable at all. The

mere gene sequence information is sufficient to have a template for the chemical synthesis of the

corresponding DNA that can be subsequently inserted into an appropriate expression system.

Heterologous expression has thus lead to the expression of highly interesting NPs from a variety of

remarkable microorganisms that thrive under extreme conditions or are otherwise difficult to

cultivate. Examples include: purine nucleoside phosphorylase from the cold- adapted marine

bacterium Pseudoalteromonas sp. Bsi590 (Li et al. 2008), uridine phosphorylase from the pathogenic

protozoan Trypanosoma brucei (Larson et al. 2010), purine nucleoside phosphorylase from the

hyperthermophilic archaeon P. furiosus (Cacciapuoti et al. 2007), 5′-deoxy-5′-methylthioadenosine

phosphorylase from S. solfataricus (Cacciapuoti et al. 2005), as well as the purine nucleoside

phosphorylase from the alkaliphilic microorganism Bacillus halodurans (Visser et al. 2010). In future,

metagenomic approaches that have been recognized as powerful strategies for the discovery of new

genes within the search for “ideal” biocatalysts (Cowan et al. 2004), might also become important

tools for the discovery of novel NP variants. Recently, Cieslinski and co-workers have found a MTAP

that resembles the MTAP from Psychrobacter arcticus 273-4 by screening a metagenomic library that

was generated from environmental DNA isolated from Antarctic topsoil. Remarkably the scientists

24 Introduction

were actually screening for lipolytic enzymes; and the clone expressing the MTAP has just captured

their interest because it showed unusual degree of pink fluorescence on the screening-agar

containing rhodamine B (Cieslinski et al. 2009).

1.4. Motivation and structure of the present work

1.4.1. Enzyme-assisted synthesis of modified purine nucleosides

The ultimate goal of the present study is to improve the chemo-enzymatic route towards modified

purine nucleosides by taking advantage of novel nucleoside phosphorylases, derived from

thermophilic microorganisms. In focus is thereby the synthesis of purine nucleoside analogues with

modifications on the carbohydrate moiety (2′-fluorinated purine nucleosides) or with modifications

on the purine base (2,6-dihalogenated purine nucleosides). 2′-Fluorinated purine nucleosides display

interesting pharmaceutical activities (Bonate et al. 2006, Tuttle and Krenitsky 1992, Tuttle et al.

1993) and are furthermore valuable precursors for the synthesis of synthetic oligonucleotides with

advanced properties (Adler et al. 2008, Allerson et al. 2005, Damha et al. 1998, Deleavey et al. 2010,

Gragoudas et al. 2004, Kalota et al. 2006, Kawasaki et al. 1993, Khati et al. 2003, Manoharan et al.

2011, Morrissey et al. 2005, Ng and Adamis 2006, Watts and Damha 2008), whereas

2,6-dihalogenated purine nucleosides are convenient precursors in the preparation of various

therapeutic purine nucleoside analogues (Kazimierczuk et al. 1984, Montgomery et al. 1986, Tennilä

et al. 2000).

While the chemical synthesis of numerous pyrimidine nucleoside analogues is well established, the

synthesis of equivalent purine nucleoside analogues is often rather complicate and is typically a

multistage process that is hampered by the formation of regio- and stereoisomers. The utilization of

NPs as biocatalysts offers the possibility to transfer pentofuranosyl moieties from a pyrimidine

nucleoside pentofuranosyl donor to a purine base serving as pentofuranosyl acceptor. Such

biocatalytic applications of enzymes that operate under strict stereo- and regioselectivity is generally

a promising approach towards the development of efficient and high yielding synthetic routes, and

offers, furthermore, the possibility to design environmentally friendly processes.

By employing chemically prepared, artificial precursors as substrates, novel, modified nucleosides

can be synthesized with this strategy. Unfortunately, a number of highly interesting nucleoside

analogues are hardly recognized as substrate by NPs that are currently in use. For example the

catalytic rates of E. coli NPs in reactions involving 2′-fluorinated nucleosides are dramatically reduced

in comparison to the activities towards natural nucleosides. Therefore, high amounts of enzyme

loading and prolonged incubation times are needed to obtain reasonable yields (see e.g. Tuttle and

Krenitsky 1992). It is therefore appealing to study novel NP variants that possibly display more

favourable substrate specificities.

Introduction 25

1.4.2. Taking advantage of thermostable nucleoside phosphorylases

A major limitation towards the wide application of enzymes in chemical synthesis concerns the

instability under harsh reaction conditions, for example at high temperature. However, the operation

at higher temperature has many advantages: higher reaction rates can be achieved, due to a

decrease in viscosity and an increase in diffusion coefficients of substrates. The final yield may be

increased due to higher solubility of substrates that are only poorly soluble at ambient temperature,

as it is also the case for many purine bases. And finally there is a lower risk of contamination by

common mesophiles.

For these reasons inherently thermostable enzymes derived from thermophilic microorganisms have

been recognized as powerful biocatalysts for industrial processes (Haki and Rakshit 2003). It has been

shown, that thermal stability is not the result of a unique phenomenon. Instead, nature has found

diverse strategies towards enzyme stabilization: increased electrostatic interactions, greater

hydrophobicity and better atom packing, deletion or shortening of loops and disulfide bonds are just

some examples. By taking advantage of enzymes from thermophiles the intrinsic stability of these

proteins, that is the product of evolution, can be exploited. Noteworthy thermal stability is often

accompanied by a higher resistance against denaturants and other harsh reaction conditions that

may be encountered in industry.

1.4.3. Recombinant expression of thermostable NPs in E. coli

Since thermophilic microorganisms are generally difficult to cultivate to high cell densities (Krahe et

al. 1996), the target enzymes of this study will be expressed in E. coli. This approach offers the

possibility to produce the enzymes in high yields and simplifies the downstream processing: Major

parts of the mesophilic endogenous E. coli enzymes can be simply removed by heat precipitation.

Although recombinant expression can be understood primarily just as a tool here, it should be noted

that the outcome is pivotal for the entire study. In fact, the efficient expression of thermostable

enzymes is challenged by various factors. Knowledge gained in this field of research may further

support the use of thermostable enzymes in industrial processes.

1.4.4. Experimental outline and key objectives

For the realization of the aims outlined above, three work packages have been defined. The results of

each work package will be presented in each one of the following chapters.

The first work package represents the Recombinant expression of thermostable NPs. This work

package includes the selection of target enzymes, cloning and optimization of the expression in

E. coli. The second work package is the Characterization of recombinant NPs. Here the aim is to

determine characteristics of the enzymes that will be important for the final application. In focus will

be temperature optima, thermal stabilities and substrate spectra of the enzymes. The final work

package is the Application in the synthesis of modified nucleosides. This work package will show

26 Introduction

whether the generated thermostable enzymes are indeed beneficial for the synthesis of modified

purine nucleosides that have been defined as target compounds.

2. Experimental part

2.1. Generation of expression plasmids

2.1.1. Gene isolation

G. thermoglucosidasius 11955 was grown for 1.5 days at 52 °C in LB (10 g l-1 tryptone, 5 g l-1 yeast

extract, 10 g l-1 NaCl, pH 7.0). The genomic DNA was isolated using an adapted standard protocol

(Moore and Dowhan 2002). Shortly the procedure involved the treatment with proteinase K in the

presence of SDS, ethanol precipitation of nucleic acids, a washing step (with 70 % ethanol) and re-

suspension of the air-dried pellet in a solution containing RNase A. Genomic DNA of T. thermophilus

HB27 was obtained from Thomas Schwab from the Institute of Biophysics and Physical Biochemistry,

University of Regensburg. Genomic DNA of D. geothermalis and A. pernix were isolated before by

Marco Casteleijn in the Bioprocess Engineering Laboratory of the University of Oulu, Finland. The

target gene sequences were amplified from genomic DNA using Pfu DNA polymerase (Fermentas,

Lithuania; now Thermo Scientific).

2.1.2. Plasmids

The pCTUT7 vector was chosen as basis for the generation of NP expression vectors. This vector is

characterized by an IPTG inducible lac promoter derivate and a pBR322 origin of replication. In

addition this vector contains attR recombination sites through which a gene coding sequence can be

conveniently introduced by recombinational cloning. For more details see (Šiurkus et al. 2010).

For some expression studies pET21a (Novagen) was used that is based on the T7 expression system.

The T7 promoter located upstream of the target sequence on the plasmid, is recognized by the RNA

polymerase of the T7 bacteriophage. The corresponding expression cassette is usually genomically

integrated in the expression strains, indicated by the genotype annotation (DE3). The T7 expression

system is considered as very efficient.

Table 2. Chaperone plasmid set (Takara)

Plasmid Chaperone Promoter Inducer

pG-KJE8 dnaK-dnaJ-grpE

groES-groEL

araB

Pzt1

L-Arabinose

Tetracyclin

pGro7 groES-groEL araB L-Arabinose

pKJE7 dnaK-dnaJ-grpE araB L-Arabinose

pG-Tf2 groES-groEL-tig Pzt1 Tetracyclin

pTf16 Tig araB L-Arabinose

28 Experimental part

For co-expression experiments a set of chaperone plasmids (Takara Bio Inc., Otsu, Japan) was used.

Chaperone expression, that has been shown to promote proper folding of recombinantly expressed

proteins (Baneyx and Mujacic 2004, Nishihara et al. 1998, Schlieker et al. 2002), is here induced by

addition of L-arabinose or tetracycline. The plasmids have a chloramphenicol resistance marker and a

p15A origin of replication. For more details see Table 2.

2.1.3. Recombinational cloning

a) Construction of the DgPNP entry vector (pEntrDgeo1497)

The following primer pair was used to amplify the gene coding for DgPNP: forward -

5′ GGGGACAAGTTTGTACAAAAAAGCAGGCTTCGAAAACCTGTATTTTCAGGGCATGGTGGTGGCGCGTGTAC

CGG 3′, reverse - 5′ GGGGACCACTTTGTACAAGAAAGCTGGGTTCACATGCTGTTGGAAGGTACT 3′ (the

underlined portion represents the gene specific template, the 5′ overhang includes attB1

recombination site and an engineered TEV cleavage site in the forward primer and the attB2

recombination site in the reverse primer). The purified PCR fragment (PeqGold gel extraction kit C-

line, Peqlab Biotechnologie GmbH, Erlangen, Germany) was inserted into the pDONR201™ vector

(Invitrogen) by the site specific “BP” recombination reaction with BP Clonase II enzyme mix

(Invitrogen) according to the manufacturer’s manual.

b) Construction of DgPNP expression vectors (pCTUT7_DgPNP and pCTUT7A_DgPNP)

Following Invitrogen’s manual, the DgPNP gene was transferred from the entry vector to the

cytoplasmic expression vector pCTUT7 (Šiurkus et al. 2010) that confers a N-terminal hexahistidine

tag, by the site specific “LR” recombination reaction (LR clonase II enzyme mix, Invitrogen). The

resulting expression vector was assigned as pCTUT7_DgPNP. The DgPNP gene was also transferred

into the pCTUT7A vector, a derivative of the pCTUT7 vector, resulting in the plasmid

pCTUT7A_DgPNP. Construction of the pCTUT7A destination vector is described in the following

section.

A) B)

Figure 10: Vector maps of destination vectors pCTUT7 (A) and pCTUT7A (B) in which the chloramphenicol resistance cassette (Cmr) was replaced by an ampicillin resistance cassette (Ampr) and the plasmid stabilizing parB locus.

Experimental part 29

2.1.4. Cloning via restriction and digestion

PCR-amplified fragments containing the gene coding sequences (see Appendix) were cloned into the

vector backbones through the treatment with FastDigest restriction endonucleases (Fermentas,

Lithuania) and subsequent ligation (T4 DNA Ligase, Roche). The oligonucleotide sequences

(synthesized by TIB Molbiol, Berlin Germany) that were used for the amplification of the inserts,

restriction enzymes and vector backbones used for the construction of each expression vector are

summarized in Table 3.The Vector NTI software (Invitrogen) was used to plan and analyze cloning

procedures.

The plasmid pCTUT7 was modified by replacing the chloramphenicol with an ampicillin resistance

cassette, and introducing the plasmid stabilizing parB locus [22]. Therefore the ampicillin resistance

cassette was amplified with the first primer pair shown Table 3 in the first line. The 580 bp long parB

region of plasmid R1 (GenBank A20060.1) was amplified with the second primer pair from the

plasmid pKG1022, kindly provided by Prof. Thomas Schweder. The resulting vector was assigned as

pCTUT7A, and was used as destination vector for recombinational coning, as well as vector backbone

for the construction of expression vectors via restriction and ligation (see Table 3).

For the construction of the plasmid pKS2_ApUP (Figure 11) the primer extensions for the insert

amplification were designed in such a way, that a hexahistidine tag is fused to the N-terminus of the

target protein. A BamHI restriction site was engineered between hexahistidine and gene coding

sequence; hence the pKS2_ApUP vector could be conveniently used for cloning of the other target

genes to be expressed with N-terminal hexahistidine tag (Table 3).

AGGAGATATACATATGATGAGAGGATCGCATCACCATCACCATCACGGATCCGGAGACGAGAGTCTA

NdeI BamHI6 x his ApUP

Figure 11: Vector map of pKS2_ApUP. Native gene sequences can be inserted in this vector by NdeI/HindIII digestion. Alternatively inserts can be cloned by BamHI/HindIII digestion; in this case a hexahistidine tag will be fused to the N-terminus of the target protein.

30 Experimental part

2.1.5. Verification of the cloning steps and propagation of vector constructs

The progress of cloning was followed by the visualization of amplified DNA fragments and digested

vectors on agarose gels. After the ligation step, plasmids were transformed by electroporation

(electroporator 2510, Eppendorf) into E. coli TOP10 (for details see Table 5) following a standard

protocol (Sambrook and Russell 2001). For the propagation of pCTUT7 and pCTUT7A an E. coli strain

with resistance to the ccdB gene product (E. coli ccdB+, Table 5), encoded on both vectors, was used.

Single colonies obtained after overnight cultivation on agar plates were screened for the presence of

the target gene sequence via PCR. Positive clones were cultivated and the plasmids were extracted

and purified (Invisorb® Spin Plasmid kit, Invitek, Berlin, Germany). The plasmid concentration was

determined by a spectroscopic measurement at 260 nm (Nanodrop, Thermo Scientific). In a next step

the plasmids were digested with restriction enzymes. If the restriction pattern was in accordance

with the theoretically expected pattern, plasmids were sent for sequencing to confirm that the target

sequence is correct.

2.1.6. Site-directed mutagenesis

Site-directed sequence optimizations were normally introduced by the primers that served for insert

amplification. With this method the following expression plasmids were generated: pKS1_DgPNP1,

pKS1_DgPNP2, pKS1_ApMTAP1, pKS1_ApMTAP2, pKS1_ApUP1 (see Table 3 for corresponding primer

pairs). Another approach involved the amplification of an entire plasmid that served as template.

With this method the GtPyNP sequence was modified to match the corresponding data bank amino

acid sequence. Kapa Hifi DNA polymerase (Kapa Biosystems, Woburn, United States) that is especially

suitable for long range amplification with high fidelity was used for this purpose. The template

sequence was afterwards removed by DpnI digestion. The freshly, in vitro synthesized plasmids

containing the modified sequence are not recognized by DpnI since they lack methylation patterns

generated in vivo by dam positive E. coli strains. The mutagenesis primers used for this purpose

were:

Forward primer: 5′ GCAGGAGCGAAGCGGCTCGCAACAGCGATG 3′

Reverse primer: 5′ CATCGCTGTTGCGAGCCGCTTCGCTCCTGC 3′

Experimental part 31

Table 3. Construction of expression vectors via restriction and digestion

Vector Primer pair for insert amplification (5′ – 3 ′) Vector backbone

Restrict. enzymes

pCTUT7A F1: GAAATGTGCGCGGAACC

R1: TACTAGCCATGGCAATCTAAAGTATATATGAGTAAAC

F2: GGTTCCGCGCACATTTCAACAAACTCCGGGAGG

R2: TGCATGAAGCTTACAACATCAGCAAGGAGA

pCTUT7 NcoI

HindIII

pKS1_DgPNP F: TACTAGCATATGGTGGTGGCGCGTGTAC

R: CAGCATAAGCTTTCACATGCTGTTGGAAGG

pCTUT7A NdeI

HIndIII

pKS1_DgPNP1 F: TACTAGCATATGATTGCGCGTGTACCGGCAA

R: same as for pKS1_DgPNP

pCTUT7A NdeI

HIndIII

pKS1_DgPNP2 F: TACTAGCATATGATTGCCCGAGTACCCGCACGTCCTTTCGCTTCCCCGC

R: same as for pKS1_DgPNP

pCTUT7A NdeI

HIndIII

pKS2_DgPNP F: TACTAGGGATCCGTGGCGCGTGTACCGG

R: same as for pKS1_DgPNP

pKS2_ApUP BamHI

HindIII

pKS1_ApMTAP F: ACTAGCATATGAGGAAGCCGGTTCACCTCG

R: AGCATAAGCTTCTAGACTCCTCCTGTGAGG

pCTUT7A NdeI

HIndIII

pKS1_ApMTAP1 F: ACTAGCATATGAGGAAGCCAGTTCACCTAGAGGCAGGGCCCGGC

R: same as for pKS1_ApMTAP

pCTUT7A NdeI

HIndIII

pKS1_ApMTAP2 F: ATATACATATGAGGAAACCAGTACACCTAGAGGCAGGGCC

R: same as for pKS1_ApMTAP

pCTUT7A NdeI

HIndIII

pKS2_ApMTAP F: ACTAGGGATCCAGGAAGCCGGTTCACCTCG

R: same as for pKS1_ApMTAP

pKS2_ApUP BamHI

HindIII

pET21a_GtPNP F: ACTAGCATATGAGCATCCATATCGAAGCAA

R: AGCATGAATTCTTATTCTACGCGAATCGCC

pET21a NdeI

EcoRI

pET21a_GtPNPh F: same as for pET21a_GtPNP

R: CAGCATAAGCTTTTCTACGCGAATCGCCG

pET21a NdeI HindIII

pKS2_GtPNP F: ACTAGGGATCCTTGAGCATCCATATCGAAG

R: AGCATAAGCTTTTATTCTACGCGAATCGCC

pKS2_ApUP BamHI HindIII

pET21a_GtPyNP F: ACTAGCATATGGTCGATTTAATTGCGAAAA

R: AGCATGAATTCTTATGAAATGGTTTCGTAT

pET21a NdeI EcoRI

pKS2_GtPyNP F: ACTAGGGATCCATGGTCGATTTAATTGCGA

R: AGCATGCGGCCGCTTATGAAATGGTTTCGTATATA

pKS2_ApUP BamHI

NotI

pKS2_TtPyNP F: ACTAGGGATCCAACCCCGTGGTCTTCATC

R: AGCATGCGGCCGCCTAGATGGCCTCCAGGA

pKS2_ApUP BamHI

NotI

pKS1_ApUP F: TACTAGCATATGGGAGACGAGAGTCTAAGG

R: CAGCATAAGCTTCTATGTGCGTCTGCACGC

pCTUT7A NdeI HindIII

pKS1_ApUP1 F1: same as for pKS1_ApUP

R1: GGGCCACATCCCCACGCCGGACCCTCAGATGG

F2: CCATCTGAGGGTCCGGCGTGGGGATGTGGCCC

R2: same as fpr pKS1_ApUP

pCTUT7A NdeI HindIII

pKS2_ApUP F1: CATCACCATCACCATCACGGATCCGGAGACGAGAGTCTAAGG

F2: TACTAGCATATGATGAGAGGATCGCATCACCATCACCATCACGG

R: same as for pKS1_ApUP

pCTUT7A NdeI HindIII

pKS2_ApUPsh F: ACTAGGGATCCGTGGCCCGCTACGTTCTCC

R: same as for pKS1_ApUP

pKS2_ApUP BamHI HindIII

Bold letters indicate restrictions sites and the underlined portion represents the insert-specific sequence. F = forward, R = reverse.

32 Experimental part

2.2. Bioinformatics

Table 4. Bioinformatic tools

Program Reference Application

Vector NTI Invitrogen Vector design,

molecular weight prediction

CYSPRED (Fariselli and Casadio 1999) Cysteine oxidation state

prediction

GenScript Rare Codon

Analysis Tool

GenScript

(https://www.genscript.com)

Codon adaption index

determination

Rare codon caltor http://people.mbi.ucla.edu/sumchan/

caltor.html

Rare codon analysis

Mfold (Zuker 2003) mRNA secondary structure

prediction

CCP4mg (McNicholas et al. 2011). Visualization of 3D protein models

2.2.1. Amino acid sequence analysis and homology modelling

Amino acid sequence identities were assessed with the protein basic local alignment tool (BLAST)

(Altschul et al. 1990) of the National Center for Biotechnology Information web server

(http://blast.ncbi.nlm.nih.gov). For multiple sequence alignments and phylogenetic analysis the

ClustalW2 - multiple sequence alignment or the ClustalW2 - phylogeny tool (Larkin et al. 2007),

respectively, have been used, both accessible through the European Bioinformatics Institute web

server (http://www.ebi.ac.uk). Three-dimensional models of the protein structures of the target

proteins were built by homology modelling using the Swiss-model workspace (Bordoli et al. 2009).

2.2.2. Secondary mRNA prediction and sequence optimization

Secondary mRNA prediction and optimization of the 5′ mRNA of the DgPNP constructs are described

in detail in (Szeker et al. 2011). For the prediction of the stability of 5′ mRNA regions of other

constructs a sequence comprising the Shine-Dalgarno sequence (AGGAGA) and the first 37

nucleotides of the coding region was used for secondary structure prediction (Kudla and co-worker

have shown, that after these 37 coding nucleotides, the correlation between mRNA folding energy

and protein expression drops significantly (Kudla et al. 2009)). For retrieving the minimum free

energy of secondary structure formation in this region, the mfold web server was used (Zuker 2003).

Optimized sequences were generated by synonymous codon substitutions.

Experimental part 33

2.3. Bacterial growth and recombinant protein expression

Table 5. Genotype of the used E. coli strains.

Strain Genotype Company

E. coli TOP10 F–mcrA Δ(mrr-hsdRMS-mcrBC) φ80lacZΔM15 ΔlacΧ74 recA1

araD139 Δ(ara-leu) 7697 galU galK rpsL (StrR) endA1 nupG λ-

Invitrogen

E. coli ccdB+ F– mcrA Δ(mrr-hsdRMS-mcrBC) Φ80lacZΔM15 ΔlacX74 recA1

araΔ139 Δ(ara-leu)7697 galU galK rpsL (StrR) endA1 nupG fhuA::IS2

Invitrogen

E. coli BL21 F– ompT hsdSB(rB– mB

–) gal dcm Novagen

E. coli Origami Δ( ara–leu)7697 ΔlacX74 ΔphoA PvuII phoR araD139 ahpC galE galK rpsL F'[lac+ lacI )pro] gor522::Tn10 trxB (KanR, StrR, TetR)

Novagen

E. coli Rosetta2 F–ompT hsdSB(rB– mB

–) gal dcm pRARE2 (CamR) Novagen

E. coli Rosetta (DE3) F–ompT hsdSB(rB– mB

–) gal dcm (DE3) pRARE2 (CamR) Novagen

2.3.1. Preparation of recombinant E. coli cell banks

The generated expression vectors (Table 3) were transformed into appropriate E. coli expression

strains (Table 5). Single colonies were picked and cultivated overnight in LB medium. Glycerol was

added to the culture broth to a final concentration of 25 % and aliquots were frozen and stored

at -80 °C.

2.3.2. Recombinant protein expression

Cultivations were performed in a shaking incubator with 2.5 cm shaking orbit (Lab-Therm LT-X,

Kühner, Basel, Schweiz), at 250 rpm for expression in 24-squarewell Deep well plates (HJ-Bioanalytik,

Mönchengladbach, Germany) with 10 ml total volume and a liquid volume of 3 ml, or 200 rpm for the

expression in Ultra Yield Flasks™ that were covered with AirOTop™ seals (Thomson Instrument

Company, USA).

Different media were used for expression studies. Initial expression studies with DgPNP constructs at

different temperatures were performed in LB medium. Therefore cells from LB agar plates grown

overnight were used to inoculate LB medium to an initial OD600 value of 0.2. Protein expression was

induced by addition of 100 µM or 1 mM IPTG after 2 h (cultivations performed at 42 °C), 2 1/2 h

(37 °C) and 3 h (30 °C). Cells were harvested 3 h after induction. For expression in TB medium

(Sambrook and Russell 2001) cells from LB agar plates grown overnight were used to inoculate the

main culture to an initial OD600 value of 0.1. Furthermore medium that is based on enzyme controlled

substrate delivery (EnPresso®, BioSilta, Finland) was used. Therefore the main culture was inoculated

to a final OD600 of 0.15 with a fresh preculture that was cultivated before on LB agar plates (8 – 10 h

at 37 °C). After overnight cultivation protein expression was induced by addition of IPTG, and at the

same time a “booster tablet” was added, according to the instructions of the manufacturer. Finally,

cells were harvested by centrifugation (16,000 ×g, 5 min, 4 °C) and the pellets were stored at –20 °C.

Chaperone co-expression experiments were performed as outlined in the manufacturer’s manual

(Takara Bio Inc., Otsu, Japan). Shortly, E. coli BL21, co-transformed with the expression plasmid and a

34 Experimental part

chaperone plasmid (Table 2), was cultivated in LB medium as described above; however, the medium

additionally contained 20 µg ml-1 chloramphenicol for chaperone plasmid selection, and

0.5 - 4 mg ml-1 L-arabinose or 1 – 10 mg ml-1 tetracycline for the induction of the respective

chaperones (Table 2). Three hours after incubation at 30 °C, target protein expression was induced

by the addition of 100 µM IPTG.

2.4. Preparation of protein samples

2.4.1. Cell disruption

In order to analyze protein expression by SDS-PAGE, cell disruption was performed in small-scale.

Therefore cell pellets that equalled the amount of 1 ml of OD600 = 5 were resuspended in 300 µl

BugBuster™ Protein Extraction Reagent (Novagen), containing 1 µl ml-1 Benzonase® (Novagen). The

suspension was afterwards incubated for 20 – 30 min on a shaking platform. Alternatively cells were

disrupted by sonication. For this purpose a cell pellet corresponding to the amount of cells in 1 ml of

OD600 = 5 was resuspended in 500 µl potassium phosphate buffer (10 mM , pH 7.0). This cell

suspension was sonified on ice using a UP200S sonicator (Hielscher Ultrasonics GmbH, Teltow,

Germany). The sonication was performed twice with 30 % power input for 3 min in 30 s intervals

using a sonotrode of 2 mm in diameter. After cell disruption soluble and insoluble protein fractions

were separated by centrifugation (20,000 × g; 15 min; 4 °C).

In order to prepare cell lysates for protein purification, cell pellets were re-suspended in NPI-10

binding buffer (50 mM NaH2PO4, 300 mM NaCl, 10 mM imidazole, pH 8.0) at 5 ml per gram wet

weight in a falcon tube. This cell suspension was sonified on ice using the UP200S sonicator. The

sonication was performed twice with 30 % power input for 3 min in 30 s intervals using a sonotrode

of 7 mm in diameter.

2.4.2. Protein purification

Cell lysates were centrifuged (20,000 × g, 15 min, 4 °C) to separate soluble from insoluble fractions.

The soluble portion of the cell lysate was heated for 15 min at 50 °C (DgPNP), 60 °C (GtPyNP), 65 °C

(GtPNP), 80 °C (TtPyNP), or 85 °C (ApMTAP). Coagulated proteins were removed by centrifugation

(20,000 × g, 15 min, 4 °C). Additionally the cell lysate was filtered using Rotilab®-sytinge filters, CME,

sterile, 0.22 µm pore size from Carl Roth (Karlsruhe Germany). The target proteins were further

purified with a Fast Protein Liquid Chromatography system (Äkta™ avant, GE healthcare) via

immobilized metal ion affinity chromatography using a 5 ml Ni-NTA Superflow cartridge (Qiagen),

following the instructions given by the manual. Shortly, the column was equilibrated with 10 column

volumes of NPI-10 buffer (see 2.4.1) and the cell lysate was applied using the system pump. Then,

the column was washed with NPI-20 buffer (50 mM NaH2PO4, 300 mM NaCl, 20 mM imidazole, pH

8.0) until the absorption signal at 280 nm returned to the baseline value. Finally the hexahistidine

tagged target protein was eluted with elution buffer NPI-250 (50 mM NaH2PO4, 300 mM NaCl,

250 mM imidazole, pH 8.0). Fractions containing the purified protein were pooled and subsequently

Experimental part 35

the excess of salt and imidazole was removed by the use of a HiPrep 26/10 Desalting column (GE

Healthcare). Therefore the column was equilibrated before with at least 5 column volumes desalting

buffer (50 mM KH2PO4, 0.15 M NaCl, pH 7.0). For more details see the instructions given by the

manufacturer. Finally, the purified protein solution was aliquoted, rapidly frozen in liquid nitrogen

and stored at -80 °C.

2.5. Protein analytics

2.5.1. SDS-PAGE analysis

Soluble, insoluble, and total protein fractions were analyzed by SDS-PAGE. For this purpose either

precast gels (NuPage® 4-12 % Bis-Tris gel, Invitrogen) or 15 % Tris-glycine SDS-polyacrylamide gels

with 5 % stacking gels were prepared (Sambrook and Russell 2001). Culture samples were normalized

based on their optical densities at 600 nm prior to cell lysis and subsequently prepared for SDS-PAGE

analysis as described in standard protocols (Sambrook and Russell 2001). Shortly, soluble or total

protein fractions were combined with 1 volume of 2 x SDS gel-loading buffer (100 mM Tris·Cl pH 6.8,

200 mM DTT, 4 % SDS, 0.2 % bromphenol blue, 20 % glycerol). For analysis of insoluble protein

fractions the pellet was resuspended in an adequate amount of 1 x SDS gel-loading buffer containing

8 M urea. Afterwards samples were incubated for 5 min at 99 °C, cooled to room temperature and 5

– 10 µl were applied for SDS-PAGE analysis. The protein molecular weight marker (#Sm0431) was

purchased from Fermentas (Vilnius, Lithuania; now Thermo Scientific). After gel electrophoresis

(approx. 90 min at 150 V), gels were washed with water and stained for approximately 1 h with a

solution containing 80 mg l-1 Coomassie Brialliant Blue G-250 and 35 mM HCl. The procedure does

not require extensive destaining, but in order to increase the contrast gels were washed overnight in

water.

2.5.2. Determination of the protein concentration

The protein concentration of the purified protein solution was determined by measuring the

absorption at 280 nm (Nanodrop, Thermo Scientific). The absorption coefficients were theoretically

calculated from the amino acid sequence (Vector NTI software, Invitrogen) and are listed in Table 6,

whereby the amino acid sequence that results from the expression in the pKS2 vector served as input

sequence. The theoretical prediction of the absorption coefficient at 280 nm of folded proteins in

water has been reported to be fairly reliable for proteins containing tryptophan residues (Pace et al.

1995). All target proteins of this study contain at least one tryptophan residue per subunit.

36 Experimental part

Table 6. Theoretically calculated absorption coefficients.

Protein A[280] of 1 mg ml-1

DgPNP 0.63

ApMTAP 0.88

GtPNP 0.88

GtPyNP 0.42

TtPyNP 0.56

2.5.3. Protein unfolding studies

Thermal denaturation of 10 µM purified TtPyNP protein dissolved in potassium phosphate (50 mM,

pH 7.5) was monitored with a Jasco J-815 circular dichroism (CD) spectrometer in a 0.1 cm cuvette by

following the loss of ellipticity at 220 nm. Unfolding was induced by raising the temperature in 0.1 °C

increments at a ramp rate of 1 °C min-1 with a Peltier-effect temperature controller. The measured

ellipticity was normalized, and the apparent melting temperature (TappM) was determined. DSC

experiments were performed with 43 µM TtPyNP protein dissolved in potassium phosphate (50 mM,

pH 7.5) by heating the samples in a CSC 5100 Nano differential scanning calorimeter with a scan rate

of 1 °C min-1. The DSC data were analyzed with the program CpCalc (version 2.1, Calorimetry Sciences

Corporation, 1995) to determine TappM at which half of the protein is unfolded. The irreversibility of

thermal denaturation precluded thermodynamic analysis of the CD and DSC unfolding traces.

2.6. Activity assays

2.6.1. Spectroscopic assay for PNP activity

In initial experiments, soluble fractions of DgPNP cell lysates were analyzed with respect to PNP

activity by a spectroscopic assay. Therefore the spectroscopic assay developed by Kalckar (Kalckar

1947) was adapted for use in 96-microwell plates. The final reaction mixture (200 µl) contained:

50 mM potassium phosphate buffer pH 7.5, 0.5 mM inosine (Carl Roth) and 0.4 U ml-1 microbial

xanthine oxidase (Sigma-Aldrich). After addition of 20 µl diluted soluble cell extract, the UV

compatible microplate containing the reaction mixtures was incubated for 1 min at 25 °C in a

temperature-controllable microplate reader (Synergy™ Mx, Biotek,Winoosky, United States), and

subsequently the change of absorption at 293 nm was recorded. Quantification of uric acid (Sigma-

Aldrich) was performed at the same temperature and in the same buffer, to correlate the change of

absorption to the liberation of uric acids during the enzymatic assy. One unit of PNP is defined as the

amount of PNP which liberates 1 µmol of uric acid from inosine, in the presence of excess of xanthine

oxidase.

Experimental part 37

2.6.2. Standard assay with purified proteins

Standard activity assays were performed in 50 mM potassium phosphate buffer, pH 7.0 containing

1 mM of uridine (GtPyNP, TtPyNP) or 1 mM inosine (DgPNP, GtPNP, ApMTAP) as substrate. After

2 min of pre-heating, 1-2 µl of diluted enzyme was added per 100 µl of reaction mixture. The

following standard reaction temperatures were defined for each protein: 55 °C (DgPNP), 60 °C

(GtPyNP), 70 °C (GtPNP), 80 °C (ApMTAP, TtPyNP). These temperatures either coincide with the

temperature optimum of the respective protein (DgPNP, GtPyNP, GtPNP) or were set to 80 °C due to

practical reasons (TtPyNP, ApMTAP). Samples were withdrawn after defined time intervals and the

reaction was immediately stopped by adding 1 vol of reaction mixture to ½ vol of 10 % trichloroacetic

acid (TCA). Precipitated proteins were removed by centrifugation (20,000 ×g, 15 min, 10 °C) and the

samples were stored at – 20 °C for later analysis by HPLC.

From the HPLC results the concentration of residual nucleoside (substrate) and liberated nucleobase

(product of the phosphorolytic cleavage) was retrieved. The substrate conversion was calculated

with the following formula:

%100(%)

nucleosidebase

baseConversion

Only conversion rates that were linear with respect to time and amount of enzyme added, were

considered for further analysis. This was usually the case if not more than 10 % of the substrate was

converted.

2.6.3. Thermal properties of the enzymes

Thermal stability

Enzyme preparations were diluted in 50 mM potassium phosphate buffer (pH 7.0). The final enzyme

concentrations were 13 µg ml-1 (GtPNP), 26 µg ml-1 (DgPNP), 46 µg ml-1 (GtPyNP, TtPyNP), and

55 µg ml-1 (ApMTAP). It is expected that at higher enzyme concentrations as eventually used in

synthetic applications later (typically 100 µg ml-1) enzymes will be even more stable, as it was also

reported for E. coli TP and E. coli PNP (Krenitsky et al. 1981).

Aliquots were incubated in 0.2 ml tubes in a thermocycler at the respective temperatures. After

defined time intervals, tubes were withdrawn and the residual activity of the incubated enzyme

solution was determined under the standard conditions outlined in section 2.6and plotted over the

reaction time. In order to determine the half life, the resulting curve was fitted (Sigma Plot)to the

decay function

tik

evv

0 and subsequently the half life was calculated 1

2/1 5.0ln

ikt.

Temperature optimum

Standard reaction mixtures were pre-heated for 2 min at the respective temperatures. Diluted

enzyme solutions were added and the reaction was stopped after 3 min.

38 Experimental part

2.6.4. Kinetic parameters

Activity tests were performed in triplicates for at least 5 different substrate concentrations spanning

0.25 to 5 times of the Michaelis-Menton constant. The initial reaction rates were plotted over the

substrate concentrations. The resulting substrate saturation curve was fitted to the hyperbolic

Michaelis-Menton function (Sigma Plot). From the resulting equation Km and Vmax were directly

retrieved, and kcat could be easily calculated. For details see(Copeland 2000).

2.6.5. Substrate screenings

Phosphorolysis rates for natural and modified nucleosides were determined for 1 mM nucleoside

substrate concentrations. 10 mM sodium phosphate buffer pH 6.5 was used for dUrd2′F and dUrd2′F,

reactions with other nucleosides were investigated in 50 mM potassium phosphate buffer. A typical

enzyme loading for the phosphorolysis assays with artificial nucleosides was 0.1 mg ml-1. Nucleosides

and manufacturers are listed in Table 7.

2.6.6. Synthetic reactions

Reactions were performed employing 2 mM pyrimidine nucleosides as pentofuranosyl donor and

1 mM purine base as pentofuranosyl acceptor. The buffer (pH 6.5 ) was either 10 mM NP buffer or

2 mM NP buffer, if dihalogenated purine bases were substrates of the reaction. The concentration of

the NPs was 0.1 mg ml-1 each, if not otherwise stated. Reactions were typically performed in 200 –

1000 µl scale in Eppendorf tubes that were incubated on a thermoshaker (Eppendorf) at 300 rpm.

After defined time intervals samples were withdrawn. Reactions were stopped by the addition of TCA

as outlined in section 2.6. Alternatively samples were instantly diluted in ice-cold buffer. The latter

strategy was used to investigate the synthesis of (acid labile) purine deoxyribosides.

2.7. HPLC analysis

The concentration of nucleosides and nucleobases was determined by following the absorption at

260 nm during HPLC analysis using a reversed phase C18 column (Gemini-Nx 5u, 150 × 4.60 mm,

Phenomenex, Torrance, United States) with the following gradient: from 97 % 20 mM ammonium

acetate and 3 % acetonitrile to 60 % 20 mM ammonium acetate and 40 % acetonitrile in 10 min.

Retention times and calibration factors that were determined under these conditions are listed in

(Table 7).

Authentic samples of the ribo- and deoxyribosides of dihalogenated purines were not available in the

course of this study. The retention times indicated in Table 7 in fact just represent the retention

times of the new product peaks observed in HPLC. In these cases the calibration factors are only

estimates derived indirectly from the decrease of substrate concentrations.

Experimental part 39

Table 7. Properties of nucleosides and nucleobases in HPLC analysis.

Compound Obtained from Retention

time [min]

Calibration factor

(area [AU] / [mM])

Uridine Sigma-Aldrich 3.2 5372

Uracil Sigma-Aldrich 2.4 4257

Thymidine Sigma-Aldrich 4.7 4610

Thymine Sigma-Aldrich 4.0 4365

1-(2-deoxy-2-fluoro-β-D-

arabinofuranosyl)uracil (dUrd2′F)

Metkinen Chemistry

(Kuusisto, Finland) 4.6 5449

2′-Deoxy-2′-fluorouridine (dUrd2′F) TCI Deutschland (Eschborn,

Germany) 4.4 5185

O2,2′-anhydro-1-(β-D-

arabinofuranosyl)uracil (anhydro-Urd) Prof. Igor A. Mikhailopulo 2.3 3388

1-(β-D-arabinofuranosyl)uracil (ara-U) Sigma-Aldrich 3.8 5324

Adenosine Carl Roth (Karlsruhe,

Germany) 5.0 7836

Adenine Carl Roth (Karlsruhe,

Germany) 4.2 6958

Inosine Carl Roth (Karlsruhe,

Germany) 4.0 4436

Hypoxanthine Sigma-Aldrich 3.0 4736

2′-Deoxy-2′-fluoroadenosine (dAdo2′F) Metkinen Chemistry

(Kuusisto, Finland) 5.8 15162

9-(2-deoxy-2-fluoro-β-D-

arabinofuranosyl)adenine (dAdo2′F)

Metkinen Chemistry

(Kuusisto, Finland) 5.7 11336

Cytidine Sigma-Aldrich 2.6 4224

Cytosine

2,6-Dichloropurine (26DCP)

Sigma-Aldrich

TCI Deutschland (Eschborn,

Germany)

2.0

7.6

3220

2386

6-Chloro-2-fluoropurine (6C2FP) TCI Deutschland (Eschborn,

Germany) 7.0 3676

2,6-Dichloropurine ribosides - 8.2 3610

2,6-Dichloropurine deoxyribosides - 8.8 3165

6-Chloro-2-fluoropurine ribosides - 7.5 3924

6-Chloro-2-fluoropurine deoxyriboside - 8.2 4038

(-) not purchased, but presumably synthesized.

3. Recombinant expression of nucleoside phosphorylases

3.1. Introduction

3.1.1. Recombinant expression of thermostable proteins in E. coli

For many objectives in recombinant protein expression E. coli is the first choice as expression host.

Extensive investigations within the last decades made this rod-shaped bacterium to a microbial cell

factory that is easy to cultivate and to genetically manipulate in laboratory-scale (Baneyx 1999). For

industrial-scale, methods of high cell-density cultivations have been developed that enable rapid and

cost-efficient production of recombinant proteins (Shiloach and Fass 2005). In the present study

E. coli will thus serve as expression host for the production of target enzymes derived from

thermophilic microorganisms. This strategy offers an additional advantage for later protein

purification, since the majority of E. coli enzymes can be easily removed by heat precipitation.

Despite impressive progresses, the high-level production of recombinant proteins in E. coli can be

challenging. A common problem is the aggregation of misfolded protein to insoluble inclusion bodies

or a poor product yield and a multitude of strategies to overcome such problems have been

developed. Generally these approaches address the expression system (expression vector,

expression strain, fusion tag, co-expression of molecular chaperones) (Esposito and Chatterjee 2006,

Sørensen and Mortensen 2005) or the expression conditions (cultivation temperature, aeration, level

of induction, media composition) (Berrow et al. 2006, Donovan et al. 1996, Krause et al. 2010, Schein

and Noteborn 1988). Moreover it is possible to use the degeneracy of the genetic code to re-design

coding sequences according to the requirements for high-level protein expression (Welch et al.

2009). The advances in synthetic biology in recent years make this approach a more and more

reasonable strategy.

Some factors appear to be especially relevant to be considered for the successful expression of

thermostable proteins in the mesophilic host E. coli. For example poor expression resulting from

codon bias between thermophilic donor and E. coli has been tackled by co-expression of rare tRNAs

or codon optimization (Wang and Zhang 2009). Disulfide bond formation, being a widespread feature

stabilizing intracellular proteins from thermophiles, can pose another obstacle (Beeby et al. 2005,

Cacciapuoti et al. 1999). Formation of disulfide bonds can be achieved in E. coli by translocating the

nascent protein to the oxidizing periplasm or by using mutant strains with a mild oxidative cytoplasm

(de Marco 2009). Co-expression of sulfhydryl oxidase that catalyses de novo disulfide bond formation

represents a novel technology and allows the expression of disulfide bond containing proteins in the

cytoplasm with intact reducing pathways (Hatahet et al. 2010, Nguyen et al. 2011). Other factors

including optimal folding temperature and the need of specific activation factors may play a pivotal

role for the functional expression of genes from thermophiles: Hence, in some cases the expression

of thermophilic enzymes has been successful in the thermophilic host T. thermophilus, while the

functional expression of the same enzymes in E. coli failed (Angelov et al. 2009, Hidalgo et al. 2004).

42 Recombinant expression of NPs

The potential requirement of a higher cultivation temperature for the soluble expression of genes

from thermophiles was also addressed by Koma and co-workers (Koma et al. 2006). By cultivation at

43 °C or even 46 °C they successfully expressed genes from thermophiles in E. coli that were only

hardly expressed in soluble form at 37 °C.

Another critical factor that might be of special relevance for thermostable protein expression is the

formation of secondary mRNA structures in the translation initiation region of mRNA. This

phenomenon has been recognized as an important determinant for poor recombinant protein

expression levels in general (de Smit and van Duin 1990, Griswold et al. 2003, Kudla et al. 2009) and

likewise the optimization of the 5′ mRNA sequence has been objective for a number of scientific

endeavours (Care et al. 2008, Cèbe and Geiser 2006, Griswold et al. 2003, Jung et al. 2010, Khan et al.

2007, Na et al. 2010, Niemitalo et al. 2005, Sadaf et al. 2008). Recent findings make us believe that

this factor has to be particularly addressed for the recombinant expression of genes from

thermophiles in mesophilic hosts: A computational study of Gu and co-workers (Gu et al. 2010) gives

insight into the natural variation of secondary mRNA structure stability within genes, genomes and

among different species. By analyzing the genomes of 340 species the authors showed that stability

of secondary mRNA structures within the coding sequence of genes is generally reduced near the

start codon. Moreover they found that among prokaryotes this effect becomes weaker with higher

optimal growth temperature. Since secondary mRNA structures are destabilized at higher

temperature there seems to be less selection pressure on the reduction of secondary 5′ mRNA

structure stability in thermophiles than in mesophiles. In practice this may implicate that the

expression of genes from thermophiles in mesophilic hosts, at “unnatural” low temperature, is in

particular hampered by translation inhibiting 5′ mRNA structures. For these reasons mRNA stability

will also be addressed in the present study in order to optimize the expression level of the

thermostable target enzymes.

3.1.2. Target enzymes of this study

Two main criteria were decisive for the selection of the thermostable nucleoside phosphorylases

studied in this work. On the one hand the final set of enzymes should comprise biocatalysts suitable

for the reversible phosphorolysis of both pyrimidine and purine nucleosides. The motivation lies in

the combined use of both categories of enzymes in order to carry out transglycosylation reactions as

outlined in section 1.2.3. Hence, half of the selected enzymes (Figure 12) are expected to show

substrate specificities for pyrimidine nucleosides (UP, PyNP), while the other half are expected to

show substrate specificity for purine nucleosides (PNP, MTAP).

On the other hand, the selection should represent a preferably high diversity of biocatalytical

properties in order to cover broad future application fields. In this respect the expected temperature

optima of the biocatalysts are of special interest. As mentioned before thermal stability is generally

considered as advantageous for bio-catalyzed synthesis of nucleosides and therefore in focus of this

study. However, the degree of thermal stability that is appropriate may substantially depend on the

specific process. If, for example, the temperature is restricted due to thermal lability of reactants, a

Recombinant expression of NPs 43

biocatalyst with respective moderate temperature optimum might be a better choice than a

biocatalyst exhibiting an unusually high degree of thermal stability but showing at the same time

only poor activity at the process temperature.

In order to meet these requirements, the six nucleoside phosphorylases in focus of this study

originate from different thermophilic microorganisms with distinct temperature optima (Figure 12).

To some extent, this strategy also allows studying enzymes that are formally equal but evolved in a

distinct phylogenetic context. Possibly resulting differences in substrate specificity are of high

interest for biocatalytical applications for the synthesis of modified nucleosides.

40

50

60

70

80

90

100

Tem

pop

to

f sou

rce

mic

roo

rgan

ism

[C

]

DgPNP

GtPNP

ApUP

TtPyNP

Deinococcus geothermalis

• Isolated from hot springs in Naples, Italy

• Extremely gamma-radiation resistant

• Gram-positive

Geobacillus thermoglucosidasius

• Isolated in 1983 by Suzuki and co-workers in Japan, assigned as Bacillus thermoglucosidasius

• Re-classified as Geobacillus thermoglucosidasius in 2001

• Gram-positive

Thermus thermophilus

• Isolated in the 70’s from hot springs in Japan

• Genome sequenced in 2004

• Gram-negative

Aeropyrum pernix

• Isolated from coastal solfataric thermal vent in Kodakara-Jima Island in Japan

• Areobic hyperthermophilic crenarchaeon

• Gram-negative

• Complete genome sequence published in 1999

• Re-annotation of the genome in 2006

ApMTAP

GtPyNP

°

Figure 12: Overview of target enzymes and microorganisms that served as source for gene isolation. The assigned protein names consist out of two letters representing the source microorganism followed by the functional name of the enzyme. For further details regarding the source microorganisms see references (Ferreira et al. 1997, Henne et al. 2004, Kawarabayasi et al. 1999, Nazina et al. 2001, Oshima and Imahori 1974, Sako et al. 1996, Suzuki et al. 1983, Yamazaki et al. 2006).

44 Recombinant expression of NPs

3.2. Sequence analysis and theoretical predictions

In order to anticipate potential difficulties of the recombinant expression of the target proteins, the

coding sequences were analyzed beforehand in silico. Therefore coding sequences were retrieved

from the National Centre of Biotechnology Information (GenBank accession numbers are listed in

(Table 8) and analyzed with respect to rare codon usage and probability of disulfide bond formation.

Rare codon usage is an important determinant for the efficient expression of heterologous proteins

in E. coli. Since the genetic code is degenerate, a number of different codons can code for the same

amino acid. In E. coli the usage of such synonymous codons is especially biased in genes highly and

continuously expressed during exponential growth (Hénaut and Danchin 1996), and correlates with

the abundance of corresponding tRNAs (Dong et al. 1996, Ikemura 1981). The difference of codon

usage of heterologous target gene sequences can lead to insufficient tRNA pools which can decrease

efficiency (translational stalling, premature translation termination) and accuracy (translational

frameshifting, amino acid misincorporation) during the translation of a mRNA sequence into a

protein (Akashi 1994, Kane 1995, Kurland and Gallant 1996). In Table 8 the frequency of individual

codons rarely used in E. coli (coded less than 2 % usage for the particular amino acid in genes highly

and continuously expressed during exponential growth) in the target gene sequences of the present

study are listed. In addition, the codon adaption index (CAI), as a measure for the codon usage bias

between the heterologous target sequences and genes highly expressed in E. coli was calculated for

each gene. A value of 1.0 is considered as ideal, whereas low numbers indicate that the recombinant

genes may be only poorly expressed in the desired expression host (for details see (Sharp and Li

1987)). Both, the frequency of rare codons as well as the deduced CAI indicate a general trend:

Codon bias appears to be more critical for the expression of the NP genes derived from

microorganisms with high optimal growth temperature (> 60 °C) in comparison to that of

microorganisms with lower optimal growth temperature (< 60 °C). Adequate measures to avoid the

lack of rare tRNA pools have thus been considered and will be discussed in the following sections.

The heterologous expression of thermostable proteins in the cytoplasm of E. coli may also be

challenged by the requirement of disulfide bond formation that is in some cases needed for the

stabilization of intracellular proteins from thermophiles (Beeby et al. 2005, Cacciapuoti et al. 1999).

In order to assess the likelihood of the presence of disulfide bonds in the native proteins, the

oxidation state of the cysteine residues of the target sequences was predicted. It has been previously

shown that the sequence environment of free cysteine residues and cysteines involved in disulfide

bonds differ (Fiser et al. 1992). The online tool used here for prediction of the cysteine bonding state

(CYSPRED) discriminates disulfide bonds from free cysteines by exploiting this information through

the training of a neural network. The source data are the amino acid sequences flanking cysteines of

proteins with resolved three-dimensional structure as well as evolutionary information (Fariselli and

Casadio 1999). According to the predictions a single cysteine residue (Cys112) of ApMTAP is in a

bonding state, as well as two cysteines of ApUP (Cys227 and Cys279). In case of ApMTAP it appears

hence likely that Cys112 is involved in an intermolecular disulfide bond, while for ApUP both

intramolecular and intermolecular stabilization through disulfide bonds is possible.

Recombinant expression of NPs 45

Finally the molecular weight of the monomeric subunits was theoretically predicted and is shown in

Table 8 to aid the interpretation of SDS-PAGE gels of cultivation samples shown in the following

sections.

Table 8. Properties of the gene coding regions and predictions for the gene products

Assigned abbreviation DgPNP GtPNP GtPyNP TtPyNP ApMTAP ApUP

GenBank accession

number ABF45792 EFG53380 ZP_06809030 AAS81754.1 NP_147653 NP_148386

Rar

e co

do

n a

bu

nd

ancy

AGG (Arg) 1 - - 6 17 19

CTA (Leu) - 1 2 1 3 3

AGA (Arg) - - - - - 2

ATA (Ile) - - 1 3 7 9

CGG (Arg) 6 3 5 23 1 3

GGA (Gly) 3 8 12 2 3 4

CCC (Pro) 4 - - 14 8 6

CGA (Arg) 1 1 2 - - -

CAI 0.74 0.69 0.70 0.56 0.50 0.54

Cys

tein

e

resi

du

e -

oxi

dat

ion

stat

e

101-NB 38-NB 270-NB 103-NB 112-B 225-NB

129- NB 91-NB 129-NB 227-B

279-B

Pre

dic

ted

siz

e

of

gen

e p

rod

uct

(kD

a)

Native

protein 28.5 26.1 46.0 45.4 25.6 30.3

Expressed

from pKS2 30.0 27.6 47.6 46.9 27.0 31.7

NB = non-bonding state, B = bonding state, number of the cysteine residue applies to the native amino acid sequence, CAI= Codon adaption index

3.3. Expression of DgPNP

Purine nucleoside phosphorylase from D. geothermalis (DgPNP) was chosen as first target enzyme to

be overexpressed in E. coli. Since the natural host shows optimal growth at temperature between

45 °C – 50 °C it is expected that the correctly folded recombinant protein would be temporarily

stable in this temperature range.

3.3.1. Towards the functional expression of DgPNP

For pilot expression experiments the DgPNP encoding sequence was cloned via recombinational

cloning into the pCTUT7 expression vector (Šiurkus et al. 2010). This vector is characterized by a

46 Recombinant expression of NPs

strong ribosomal binding site and a lac promoter derivative that enables fine-tuning of the

overexpression level. In addition an N-terminal hexahistidine tag is conferred. However, DgPNP

expression with this construct occurred almost exclusively in insoluble form. Various expression

optimization attempts failed; reducing the expression temperature (30 °C, 20 °C), decreasing the

IPTG concentration, testing different media, and altering the timing of induction and harvest did not

improve the amount of soluble protein significantly (data not shown).

In other studies co-expression of chaperones has been successfully applied to prevent protein

aggregation during recombinant protein expression in E. coli (Baneyx and Mujacic 2004, Nishihara et

al. 1998, Schlieker et al. 2002). Hence, the effects of chaperone co-expression on DgPNP expression

were investigated in this study. Therefore a plasmid set encoding chaperones (Takara Bio Inc., Otsu,

Japan) was used that can be co-transformed into the expression strain. For this purpose the

pCTUT7_DgPNP vector was slightly modified, by substituting the chloramphenicol resistance cassette

by the plasmid stabilizing parB locus (Gerdes 1988) and an ampicillin resistance cassette, resulting in

pCTUT7A_DgPNP. Indeed, a significant amount of soluble DgPNP was expressed (Figure 13).

Particularly the Takara plasmids pGro7, encoding groES-groEL, and pG-KJE8, encoding groES-groEL

together with dnaK-dnaJ-grpE, were found to promote soluble expression of DgPNP. However, the

resulting soluble protein was not thermostable at 50 °C (Figure 13B), and thus we concluded that

DgPNP did not obtain its native conformation.

14

66

45

35

25

1814

66

45

35

25

18

A) (kDa) M 1 2 3 4 5 6 7 8 (kDa) M 9 10 11 12B)

Figure 13: SDS-PAGE analysis of DgPNP expressed with N-terminal fusion tag in E. coli BL21 - co-expression of chaperones (A) and stability test at 50 °C (B). Cultivations at 30°C; induction of protein expression by 100 μM IPTG, 3 h after inoculation; harvest 3 h (A) or 4.5 h (B) after induction. M: molecular weight marker; 1: soluble fraction E. coli BL21; 2: soluble (and 3: total) fraction of pCTUT7A_DgPNP E. coli BL21; 4 - 8: soluble protein fractions of pCTUT7A_DgPNP BL21 with each one chaperone plasmid: pGro7, pGKJE8, pG-Tf2, pTf16 and pKJE7; 9: total (10: soluble) fraction of pCTUT7A_DgPNP pGro7 E. coli BL21; 11: total(12: soluble) protein fraction of 10 after 10 min at 50 °C. Arrows indicate putative monomeric DgPNP position.

The insertion of the DgPNP gene into the pCTUT7 vector via recombinational cloning resulted in the

fusion of 23 additional N-terminal amino acids to the DgPNP protein. These amino acids are encoded

by attB1 recombination site, an engineered TEV cleavage site and the hexahistidine tag conferred by

the vector (Figure 14). The role of these extra N-terminal amino acids on eventual folding problems

resulting in low thermostability was investigated. The artificial N-terminus was removed by cloning

Recombinant expression of NPs 47

the DgPNP gene in the same expression vector used before via restriction digestion and ligation

(pKS1_DgPNP). The theoretically determined molecular size of the monomeric unit of the new gene

product was 28.5 kDa, which is 3.1 kDa smaller than the protein expressed by the previous constructs

pCTUT7_DgPNP and pCTUT7A_DgPNP. Subsequent expression at 30 °C resulted in completely soluble

DgPNP, albeit in a low amount. Raising the cultivation temperature to 37 °C or even 42 °C, resulted in

considerably higher expression levels (Figure 15A). The new gene product was stable at 50 °C for at

least 20 min (Figure 15B), which suggested that DgPNP was folded correctly now.

His tag attB1 Tev cleavage site DgPNP

ATGCATCACCATCACCATCACGCTAGCACAAGTTTGTACAAAAAAGCAGGCTTCGAAAACCTGTATTTTCAGGGCATGGTGGTG…

Figure 14: Vector map of pCTUT7_DgPNP. Upstream of the DgPNP coding sequence, additional amino acids are encoded by attB1 recombination site, an engineered TEV cleavage site and the hexahistidine tag. .

14

66

45

35

25

18

25

14

66

45

35

18

(kDa) M30 °C 37 °C 42 °C

T S T S T S(kDa) M

0 min 10 min 20 min

T S T S T S

A) B)

Figure 15: SDS-PAGE analysis of DgPNP expressed without fusion tag at different temperatures (A), and stability test at 50 °C (B). The cultivation temperatures varied between 30 °C and 42 °C; cultivation in LB medium; 100 µM IPTG. For the stability test crude cell extract of the expression trial at 42 °C was incubated for indicated time periods at 50 °C. The arrows indicate the bands corresponding to the molecular weight theoretically calculated for the native monomeric DgPNP subunit (28.5 kDa). T = total protein fraction; S = soluble protein fraction, M = molecular weight marker.

48 Recombinant expression of NPs

3.3.2. DgPNP expression optimization by reducing secondary 5′mRNA stability

The previous section describes how a suitable expression vector was found to express DgPNP in

soluble and presumably correctly folded form. However, the poor yield of recombinant protein

prompted us to investigate strategies to increase the expression level. This section deals with the

expression optimization through the reduction of the stability of secondary mRNA structures in the

5′ region. The results presented in this section were previously published (Szeker et al. 2011). The

theoretical studies on mRNA secondary structure prediction and optimization were provided by Olli

Niemitalo and André H. Juffer (Biocenter Oulu and Department of Biochemistry, University of Oulu,

Finland).

Among other factors, the formation of secondary mRNA structures in the translation initiation region

of mRNA has been recognized as an important determinant for poor recombinant protein expression

levels in general (de Smit and van Duin 1990, Griswold et al. 2003, Kudla et al. 2009). If stable

secondary structures are formed in this (5′ mRNA) region, an efficient translation initiation is

impeded. Accordingly, the optimization of 5′ mRNA sequences to tackle this problem has been

objective for a number of scientific endeavours (Care et al. 2008, Cèbe and Geiser 2006, Griswold et

al. 2003, Jung et al. 2010, Khan et al. 2007, Na et al. 2010, Niemitalo et al. 2005, Sadaf et al. 2008).

We assumed that secondary 5′ mRNA structures are also the reason for the low expression of DgPNP,

since these structures are destabilized at higher temperature – which would explain, why the

expression level increased with higher growth temperature as shown in (Figure 15A) in the previous

section. In order to test this hypothesis the secondary-structural properties of pKS1_DgPNP mRNA

were predicted theoretically (Table 9). The free energy of formation of secondary structures in the

translation initiation region surpassed the threshold for inhibition of translation initiation of about

-6 kcal mol-1 (de Smit and van Duin 1990, de Smit and van Duin 1994) by -12 kcal mol-1 or more

depending on temperature. The Shine-Dalgarno sequence and the initiation codon were involved in

base pairs in the minimum free energy substructure encompassing the translation initiation region

(Figure 16A).

A) B) C) D)

Figure 16: Substructures of predicted free energy secondary mRNA structures.Taken from (Szeker et al. 2011).

Recombinant expression of NPs 49

With these findings the hypothesis that the formation of stable 5′ mRNA structure is causing poor

expression levels in the pKS1_DgPNP construct was further supported. Hence, our aim was to

attenuate the stability of these structures. Our strategy to follow this purpose was to combine the

following two approaches:

(i) Optimization of the 5′ mRNA sequence

(ii) High-temperature cultivation

Design of optimized 5′ mRNA sequences

The gene sequence was optimized in silico to reduce secondary structures. Using the pKS1_DgPNP

mRNA as template, the second codon, GTG, was removed, as its function as a start codon had been

superseded by the artificially introduced ATG start codon, and because in the second codon position

the originally N-formylmethionine-encoding GTG encoded for a nonnative valine. It was also thought

that elimination of the G-rich codon might reduce formation of secondary structures (Kudla et al.

2009). However, no silent codon substitutions could be found by automated optimization that would

have reduced the stability of secondary structures to a satisfactory level. This was the case both

before and after removal of the GTG. Amino-acid-changing codon substitutions were thus

considered, using as template the sequence from which the GTG at the second codon position had

been deleted. The new second codon was also GTG, encoding for valine. To avoid affecting the

protein structure too much, only codons encoding for physico-chemically similar amino acids were

considered in place of the valine, and ATT, encoding for isoleucine, the most similar (Wei et al. 1997)

amino acid, was chosen. The substitution was implemented along with the earlier removal of GTG,

forming DgPNP1. For pKS1_DgPNP1, the free energy of secondary structure formation in the

translation initiation region was still at least -3.7 kcal mol-1 lower than the threshold for inhibition. On

the other hand, the secondary structures were more dynamical (Table 9), not adequately described

by the minimum free energy substructure (Figure 16B).

With DgPNP1 as the template three additional silent substitutions, Ala3: GCGGCC, Arg4:

CGTCGA, Pro6: CCGCCC were found by an automated optimization. Those together with the

silent substitution Arg8: AGGCGT resulted in DgPNP2. The free energy of secondary structure

formation at the translation initiation region of pKS1_ DgPNP2 was at 42 °C approximately at the

threshold for uninhibited translation initiation. The minimum free energy substructure (Figure 16C,D)

had little value in describing the largely unordered secondary structures (Table 9). For all three gene

variants, the free energy of secondary structure formation was lower for lower temperatures, with a

difference of at least 3 kcal mol-1 between 30 °C and 42 °C and at least 1.3 kcal mol-1 between 37 °C

and 42 °C.

Protein expression and activity tests of derived constructs

The effects of the 5′ mRNA optimizations and high-temperature cultivation on the expression of

functional DgPNP were analyzed in parallel. Therefore samples of E. coli strains expressing the

50 Recombinant expression of NPs

different DgPNP constructs at temperatures between 30 °C and 42 °C were analyzed by denaturing

gel electrophoresis and an activity assay.

The presence of E. coli intrinsic PNP was a critical issue for the determination of DgPNP activity from

crude soluble E. coli extracts. However, activity tests of E. coli BL21 cells alone and E. coli BL21 cells

expressing the recombinant enzyme revealed a sufficiently low E. coli PNP background (Figure 17A).

In the following experiments PNP activities of E. coli BL21, cultivated under same conditions as the

recombinant cultures, were determined and served as blank values.

Table 9. Secondary-structural properties of mRNA substructures of DgPNP gene variants at the translation initiation region. Adapted from (Szeker et al. 2011).

Variant 5′codons T (°C) ΔGF

(kcal mol-1

)

Substructure Base pairs

(bp)

Mean identity

(bp)

DgPNP ATG-GTG-GTG-GCG-CGT-GTA-CCG-GCA-AGG 30 -22.6 Figure 16A 17 16.1

37 -19.9 15.7

42 -17.9 15.3

DgPNP1 ATG-ATT-GCG-CGT-GTA-CCG-GCA-AGG 30 -13.4 Figure 16B 12 6.2

37 -11.3 6.8

42 -9.7 7.0

DgPNP2 ATG-ATT-GCC- CGA-GTA-CCC-GCA-CGT 30 -9.6 Figure 16C 8 3.5

37 -7.7 3.1

42 -6.4 Figure 16D 14 5.7

Compared to the original sequence, variants DgPNP1 and DgPNP2 lacked the second codon and contained a number of nucleotide substitutions (bold). The free energy of secondary structure formation (ΔGF) at the translation initiation region was predicted. The substructure of the minimum free energy secondary structure concerning the translation initiation region was found to be identical at different temperatures for DgPNP and DgPNP1, but differed between 30/37 °C and 42 °C for DgPNP2. To see how well each substructure described the dynamical secondary structures at the translation initiation region, the number of base pairs in the substructure was calculated to be compared against the Boltzmann-weighted mean number of shared base pairs between it and ensemble structures (mean identity). The presence of competing structures drops mean identity below the number of base pairs in the structure.

Optimization of the 5′ mRNA enhances total DgPNP expression

The level of total protein expression correlates well with the predicted free energies associated with

5′ mRNA folding (Table 9 and Figure 17B). The highest free energy, for example, was predicted for the

5′ mRNA folding of the DgPNP2 construct. The almost complete elimination of translation initiation

inhibition by secondary 5′ mRNA structure formation was thus expected. In agreement with this

prediction a very high amount of total DgPNP2 expression was detected under all conditions tested.

The soluble expression of DgPNP and the exerted PNP activity were significantly increased in the

constructs with the optimized 5′ sequence (Figure 17C), too. Nevertheless, SDS-PAGE analysis (Figure

17B) clearly shows that the level of total DgPNP expression was more enhanced than the level of

soluble DgPNP expression. This implies that the 5′ mRNA optimization provoked to some extent the

formation of insoluble protein. The only exception is the expression trial of DgPNP1 at 30 °C, here

5′ mRNA optimization resulted in equal higher levels of soluble and total DgPNP expression.

Recombinant expression of NPs 51

Various strategies of 5′ mRNA optimization for the enhancement of recombinant protein expression

have been reported before (Care et al. 2008, Cèbe and Geiser 2006, Griswold et al. 2003, Jung et al.

2010, Khan et al. 2007, Na et al. 2010, Niemitalo et al. 2005, Sadaf et al. 2008). The strategy

presented here, demonstrates the benefits of a computational approach.

Higher cultivation temperature increases the yield of soluble and active DgPNP

Raising the cultivation temperature from 30 °C to 37 °C or even 42 °C resulted in considerably higher

yields of soluble and active recombinant DgPNP. This applied both to the original pKS1_DgPNP

construct as well as to the new pKS1_DgPNP1 and pKS1_DgPNP2 constructs (Figure 17B, C). For

example the DgPNP activity per cell harbouring pKS1_DgPNP resulted in an over 10-fold increase by

increasing the cultivation temperature from 30 °C to 42 °C.

This phenomenon can be well explained by the temperature dependence of secondary structure

stability. If sufficient heat is supplied to compensate for the base bonding energies, secondary RNA

structures are disassembled. Likewise translation inhibiting 5′ mRNA structures (de Smit and van Duin

1990, Griswold et al. 2003) are destabilized or completely eliminated at elevated temperatures and

translation initiation efficiency is increased. The same concept of translational control by heat is

known from so-called RNA thermometers. Through secondary structure formation of mRNA,

translation is inhibited at lower temperatures. At higher temperature the structure melts and

translation initiation can occur. Such RNA-based genetic control systems can be found in heat- and

cold-shock responses or in virulence gene expression (Narberhaus et al. 2006). Neupert and co-

workers exploited this principle for temperature-controlled recombinant protein expression in E. coli

and designed synthetic RNA thermometers (Neupert et al. 2008).

High-temperature cultivation versus 5′ mRNA optimization

The experimental design allows a separate analysis of the effects of high-temperature cultivation and

5′ mRNA optimization. SDS-PAGE analysis of total protein fractions revealed that the 5′ sequence

optimization had a stronger impact on total protein expression than the high-temperature approach

(Figure 17B). Regarding the increase of exhibited PNP activity per cell, however, the high-

temperature cultivation approach seems to be more effective than the optimization of the 5′ mRNA

(Figure 17C). Hence, other factors than the stability of 5′ mRNA structures and their curtailing effect

on total protein expression have to be considered for the interpretation of the results.

A possible cause for the described phenomenon lies in the amino acid change that was introduced in

the course of 5′ mRNA optimization. One could argue that the replacement of valine by isoleucine

might lead to misfolding and aggregation, or impeded activity of the enzyme in another way. We

consider this explanation as rather unlikely for two reasons: Firstly, samples represented by a similar

intensity of soluble band also exhibited PNP activities of similar range. For example the DgPNP

construct cultivated at 42 °C displays a soluble band intensity similar to that of DgPNP1 cultivated at

37 °C. At the same time the determined activities are in a similar range (Figure 17B,C). Secondly, the

formation of insoluble protein does not seem to correlate to the new derived sequences themselves

52 Recombinant expression of NPs

but rather to the overall expression level: At 30 °C no aggregation can be observed for DgPNP1,

whereas at 37 °C and 42 °C a considerable amount of protein becomes insoluble.

T S T S T S T S T S T S T S T S T S T S T S T S T S T S T S T S T S T S

DgPNP

IPTG [mM]

Cultivation temperature

0.00

0.02

0.04

0.06

0.08

0.10

0 1 2 3 4 5 6 7

ΔA

29

3 [m

in-1

]

Soluble cell extract [%]

BL21-DgPNP2

BL21

A)

B)

C)

0.00

0.05

0.10

0.15

0.20

0.25

0.1 1 0.1 1 0.1 1 0.1 1 0.1 1 0.1 1 0.1 1 0.1 1 0.1 1

DgPNP DgPNP1 DgPNP2 DgPNP DgPNP1 DgPNP2 DgPNP DgPNP1 DgPNP2

30 C 37 C 42 C

DgP

NP

act

ivit

y [U

OD

600-1

]

° ° °

Figure 17: Expression analysis of cultures expressing the original and the optimized DgPNP constructs at different temperatures: E. coli BL21 PNP background activity (A), SDS-PAGE analysis of protein fractions from OD600-normalized samples (B), and determined recombinant PNP activities (C). The PNP activities of soluble protein fractions from E. coli BL21 alone (BL21) and E. coli BL21 expressing DgPNP2 (both at 42 °C) were analyzed with respect to the percentage of crude cell extract in the sample (A). The same cultivation samples from DgPNP, DgPNP1 and DgPNP2, expressed at 30 °C, 37 °C, or 42 °C and induced with 100 μM or 1 mM IPTG served for SDS-PAGE analysis (B) and for the determination of DgPNP activity (C). T= total; S= soluble protein fractions. Error bars: standard deviation from 3 independent enzymatic reactions. Figure adapted from (Szeker et al. 2011).

Other interpretations of the beneficial effects of the high-temperature cultivation beyond the

destabilization of translation inhibiting 5′ mRNA might be found in the scientific literature. Hence, the

transient exposure of E. coli to 42 °C, known as heat shock treatment, has been shown to enhance

Recombinant expression of NPs 53

the soluble expression of some recombinant proteins (Chen et al. 2002, Oganesyan et al. 2007). The

authors discuss the potential role of molecular chaperones, which are part of the heat shock

response and can assist folding or prevent protein aggregation. Another possible explanation for the

beneficial effect of the high-temperature cultivation on enzyme activity may be found in the

thermophilic nature of the purine nucleoside phosphorylase expressed here. Thermophilic proteins

were found to be characterized by slow folding rates (Kaushik et al. 2002, Ogasahara et al. 1998)

which should increase with higher temperature. Indeed it was shown that the solubility of a number

of thermophilic proteins was enhanced when cultivating the recombinant E. coli strains at high

temperatures (Koma et al. 2006). In the same study chaperone co-expression experiments were also

performed and did not improve the expression, making it unlikely that increased levels of molecular

chaperons are responsible for the positive result.

Combination of both methods yields best result

The combination of sequence optimization and high-temperature cultivation resulted in the highest

PNP activity per cell (Figure 17C). Compared to the original sequence expressed at 30 °C the activity

per cell was increased over 18-fold by expressing optimized PNP sequences at 42 °C.

3.3.3. Functional expression of DgPNP with N-terminal hexahistidine tag

The preceding two subsections have shown that i) DgPNP is not correctly folded if expressed with

“long” N-terminal tag and that ii) secondary structures within the 5′ mRNA impede an efficient

protein expression if this tag is removed and only the wild type DgPNP gene is expressed. Although

we successfully overcame this problem by high-temperature cultivation and optimization of the

5′ mRNA sequence, the expression of hexahistidine-tagged DgPNP would be desirable for certain

reasons. Firstly, a hexahistidine tag permits a simple and fast purification of the target protein via

immobilized metal ion affinity chromatography. Secondly, cloning a hexahistidine coding tag

upstream of the target gene provides a defined 5′ mRNA sequence in the region relevant for an

efficient translation initiation. If one time shown that the expression works well with such a

construct, a variety of target genes could be cloned downstream of this sequence without the need

to bother about secondary 5′ mRNA structure formation and the destabilization thereof for protein

expression. Having in mind that the aim is to clone and express 5 more target enzymes (for details

see section3.1.2), it would be hence of great value to make use of such straight-forward cloning

approach. This motivated us to investigate, whether the expression of correctly folded DgPNP with

N-terminal his tag is possible, despite the negative results shown for the pCTUT7_DgPNP construct in

section 3.3.1. Since we assumed that the multitude of additional amino acids at the N-terminus of

DgPNP lead to folding problems in this original vector construct, the aim was now to remove the

superfluous codons between N-terminal histidine tag and the start of the target gene sequence. The

resulting vector will be assigned as pKS2_DgPNP.

Initial expression experiments with pKS2_DgPNP were performed in TB medium at 37 °C. As in the

previous experiments, the E. coli BL21 strain was used. Protein expression was induced by adding

IPTG to a final concentration of 50 µM. Under these conditions, DgPNP was mostly expressed in

54 Recombinant expression of NPs

insoluble form (Figure 18). However, in contrast to the previous experiments with pCTUT7_DgPNP

(Figure 13), a significant amount of soluble DgPNP can be found, too. In addition, the amount of

soluble DgPNP does not decrease after thermo-treatment at 50 °C which indicates that DgPNP is

correctly folded.

14

18

25

35

45

66

0

4

8

12

16

20

0 1 2 3 4 5 6O

pti

cal d

ensi

ty (

60

0 n

m)

Cultivation time [h]

Induction

(kDa) M

Before 50 °C incubation

After 50 °C incubation

S IN S IN

A) B)

Figure 18: Expression of DgPNP with N-terminal hexahistidine tag from E. coli BL21 pKS2_DgPNP. (A) SDS-PAGE analysis of soluble (S) and insoluble (IN) protein fractions before and after a 5 min incubation step at 50 °C. (B) Optical density at 600 nm over cultivation time. Cells were cultivated in TB medium at 37 °C; 50 µM IPTG 2.5 h after inoculation. The arrow indicates the bands corresponding to the molecular weight theoretically calculated for the monomeric DgPNP subunit including N-terminal hexahistidine tag (30 kDa). M= molecular weight marker.

Since it has been shown that media with enzyme based glucose delivery have beneficial effects on

protein expression and cell density (Krause et al. 2010, Panula-Perälä et al. 2008, Ukkonen et al.

2011), EnPresso medium was used in the following experiments aiming at the optimization of the

yield of soluble DgPNP. Culture samples were analyzed by SDS-PAGE with respect to the recombinant

expression of DgPNP. Initial screening experiments showed that in samples taken 24 h after induction

significantly higher levels of recombinant DgPNP were found than in samples taken 7 h after

induction (data not shown). In the following experiments DgPNP was thus expressed for 24 h, which

is also in agreement with the instructions given by the EnPresso’s manual. Analysis of different IPTG

concentrations at 30 °C showed that 20 µM was sufficient to ensure an efficient expression of

DgPNP, while increasing the IPTG concentration to 100 µM leads mainly to a higher level of insoluble

aggregates (Figure 19A). In a next experiment the IPTG concentration was kept at 20 µM while three

different cultivation temperatures were tested (30 °C, 37 °C, 42 °C). At 42 °C significantly less soluble

protein was found than at 30 ° and 37 °C, while the amount of soluble DgPNP appeared to be slightly

higher at 37 °C than at 30 °C (Figure 19B). According to these results the optimal conditions for

DgPNP expression were defined (37 °C, 20 µM IPTG, 24 h expression) for future production of DgPNP.

Under these conditions we also obtained a reasonable cell density of OD600 = 25 (Figure 19C).

Recombinant expression of NPs 55

18

25

35

45

66

116

(kDa) M0 µM 20 µM 100 µM

IN S IN S IN S

30 °C 37 °C 42 °C

S S S

A) B)

0

5

10

15

20

25

0 10 20 30 40

Op

tica

l den

sity

(6

00

nm

)

Cultivation time [h]

C)

Induction

Figure 19: Expression study with E. coli BL21 pKS2 _DgPNP using EnPresso medium. SDS-PAGE analysis of expression study with (A) variable IPTG concentration at 30 °C; and (B) variable cultivation temperatures where cultures were induced with 20 µM IPTG. Cells were harvested 24 h after induction. (C) Growth curve of the cultivation performed at 37 °C with 20 µM IPTG. Arrows indicate the bands corresponding to the molecular weight theoretically calculated for the monomeric DgPNP subunit including a hexahistidine tag (30 kDa). M=molecular weight marker; IN=insoluble protein fraction; S=soluble protein fraction.

3.3.4. DgPNP expression - summary and conclusions

The experiments described in this section demonstrate how DgPNP can be efficiently expressed in

correctly folded form. For initial experiments the target sequence was cloned into a vector that

confers a hexahistidine tag at the N-terminus. The artificial N-terminal tag that was generated

contained not only this hexahistidine tag, but also a number of other amino acids, that were the

result of the recombinational cloning approach used. In this vector construct, DgPNP could only be

expressed in insoluble form. Chaperone co-expression appeared to aid soluble expression of DgPNP,

but thermo-treatment at 50 °C indicated that the protein had not obtained its natural conformation.

The removal of the N-terminal tag allowed the expression of soluble and active enzyme, but in

extremely low amount. The formation of translation inhibiting secondary 5′ mRNA structures was

found to be the reason for the poor expression. We tackled this problem by sequence optimization of

the 5′ mRNA and a high-temperature cultivation approach. Both methods were very effective to

reduce the stability of the 5′ mRNA and increase the yield of active DgPNP. The combination of both

methods gave the best result. This strategy might be also useful for the expression of other wild type

genes (lacking artificial N-terminal tags), especially for those derived from thermophiles. Two reasons

strongly support this assumption: The first reason is that secondary structures in the 5′ mRNA were

found to be less reduced in thermophiles than in mesophiles (Gu et al. 2010), suggesting that the

expression in mesophilic hosts is frequently impeded by 5′ mRNA structure formation. The second

reason is that the high-temperature cultivation was reported to have a positive effect on the soluble

expression of thermophilic proteins (Koma et al. 2006), which makes these proteins, in contrast to

their mesophilic counterparts, especially amenable to a high-temperature cultivation approach. In

section 3.4.1 the successful application of this strategy for the expression optimization of another

56 Recombinant expression of NPs

thermostable nucleoside phosphorylase (5′ methythioadenosine phosphorylase from the

hyperthermophilic archaeon A. pernix) will be presented.

This section has furthermore revealed that eliminating the N-terminal tag completely is not

necessary to obtain functional DgPNP. Instead, it was sufficient to reduce the N-terminal tag to the

hexahistidine sequence by cloning via restriction and digestion. This methodology opens the door for

simple purification methods based on the IMAC technology. A positive side effect is that the mRNA in

the region relevant for translation initiation consists now out of the hexahistidine coding sequence

instead of the DgPNP insert. This makes 5′ mRNA optimization and high-temperature cultivation

unnecessary for the efficient expression of DgPNP. For these reasons the resulting vector backbone

(pKS2) was also used for the expression of the other target sequences of this study.

Finally we have shown here that the yield of soluble DgPNP (when expressed from pKS2 in

E. coli BL21) could be substantially increased - mainly by minimizing the formation of insoluble

DgPNP aggregates through reduced protein expression rates. This effect was achieved by controlling

the substrate feed (EnPresso medium) and reducing the IPTG concentration.

3.4. Expression of ApMTAP

The gene coding for 5′-methylthioadenosine phosphorylase was isolated from the hyperthermophilic

archaeon A. pernix that shows optimal growth between 90 °C and 95 °C (Sako et al. 1996). It was

hence anticipated that the resulting recombinant protein (ApMTAP) would also be temporarily stable

in this temperature range.

3.4.1. Expression of the wild type ApMTAP gene without tag

The initial aim was to express the natural amino acid sequence of ApMTAP without any artificial

fusion tag. Therefore the pKS1 vector (see section3.3.1 for details) and the natural ApMTAP gene

sequence were used, the latter only modified in its start codon that was changed from TTG to ATG.

Both start codons encode N-formylmethionine, but ATG is more efficiently recognized in E. coli.

Before starting with the cloning procedure, the secondary structural properties of the resulting

pKS1_ApMTAPmRNA were predicted theoretically (Table 10). The result showed that a stable

substructure in the 5′ region is formed, that was expected to disturb efficient expression of ApMTAP

in a similar manner as discussed for DgPNP. Since the combination of 5′ mRNA sequence optimization

and high-temperature cultivation improved the expression level of DgPNP significantly, we

considered the same approach right from the start of ApMTAP expression studies. On first sight,

sequence optimization appeared to be easier in this case: two silent substitutions were found by

automated optimization (Pro4: CCGCCA and Leu7: CTCCTA). In the resulting ApMTAP1 construct

the fee energy of 5′ mRNA secondary structure formation was increased by approx. 3.6 kcal mol-1

(Table 10). SDS-PAGE analysis of culture samples showed that these substitutions alone were not

enough to achieve a considerable positive effect on the ApMTAP expression level (Figure 20A). On

the other hand, the expression level increased significantly with the expression temperature, and we

Recombinant expression of NPs 57

concluded that a better result could be obtained if secondary structures could be further destabilized

by sequence optimization. For this purpose two additional silent substitutions were introduced (Lys3:

AAGAAA, Val5: GTTGTA), resulting in ApMTAP2 (Table 10). With this optimized construct

considerably higher total and soluble expression levels could be achieved than with the original

ApMTAP sequence (Figure 20B). The same expression study was also repeated with the E. coli

Rosetta strain (Novagen) to compensate for the abundance of rare codons in the ApMTAP gene

sequence (Table 8) by overexpression of rare tRNAs. However, the resulting ApMTAP expression

patterns (Figure 20C) did not change significantly compared to the patterns obtained

withE. coli BL21.

The conclusions that can be drawn on the level of SDS-PAGE analysis are the same as described for

DgPNP expression: both sequence optimization and high-temperature cultivation lead to increased

total and soluble expression levels of ApMTAP and the best result was obtained when both methods

were combined, i.e. expression of the optimized sequence at 42 °C.

Table 10. Secondary-structural properties of ApMTAP 5′ gene variants

Variant 5′ codons ΔGF(kcal mol-1)

ApMTAP ATG-AGG-AAG-CCG-GTT-CAC-CTC -16.0

ApMTAP1 ATG-AGG-AAG-CCA-GTT-CAC-CTA -12.4

ApMTAP2 ATG-AGG-AAA-CCA-GTA-CAC-CTA -10.7

Silent nucleotide substitutions (bold) lead to variants ApMTAP1 and ApMTAP2.

3.4.2. ApMTAP expression with N-terminal hexahistidine tag

Despite the positive results presented in the previous section, the fusion of a hexahistidine tag to

ApMTAP would be of advantage, too. As discussed before, this strategy permits a simple, straight-

forward purification procedure that is a prerequisite to rapidly move forward to the characterization

of the recombinant protein. The experiments on DgPNP expression have shown that functional

expression with hexahistidine tag is possible, if the tag does not contain too many other additional

amino acids. Hence, the same vector backbone that was suitable for DgPNP expression with N-

terminal hexahistidine tag was used for the following studies with ApMTAP, resulting in the vector

pKS2_ApMTAP.

Initial screening in TB medium

The first aim was to quickly decide which expression strain is beneficial for ApMTAP expression.

Therefore both E. coli BL21 and E. coli Rosetta were investigated as expressions strains. The latter

was considered to be of advantage for ApMTAP expression, since the native gene contains numerous

codons that are only rarely used in E. coli (Table 8).

58 Recombinant expression of NPs

(kDa) M

ApMTAP ApMTAP1

30 °C 37 °C 42 °C 30 °C 37 °C 42 °C

T S T S T S T S T S T S

116

66

45

35

25

18

14

(kDa) M

30 °C 37 °C 42 °C

ApMTAP ApMTAP2 ApMTAP ApMTAP2 ApMTAP ApMTAP2

T S T S T S T S T S T S

116

66

45

35

25

18

14

A)

30 °C 37 °C 42 °C

ApMTAP ApMTAP2 ApMTAP ApMTAP2 ApMTAP ApMTAP2

T S T S T S T S T S T S

116

66

45

35

25

18

14

(kDa) M

B)

C)

Figure 20: SDS-PAGE analysis of expression studies with original (ApMTAP) and optimized (AMTAP1 and ApMTAP2) ApMTAP variants expressed in E. coli BL21 (A,B) or E. coli Rosetta (C). Cultivations were performed at 30 °C, 37 °C, or 42 °C. The cell lysates were incubated for 10 min at 90 °C prior to separation of soluble from insoluble protein fractions. Arrows indicate the bands corresponding to the molecular weight theoretically calculated for the monomeric ApMTAP subunit (25.6 kDa). T= total, S= soluble protein fractions; M= molecular weight marker.

Recombinant expression of NPs 59

The expression was performed in TB medium at 37 °C with an induction of 20 µM IPTG. Since the

gene product was expected to be highly thermostable, the culture samples were thermo-treated

at85 °C prior to SDS-PAGE analysis. The results show that a thermostable protein was expressed –

with molecular weight according to the theoretically calculated value for the monomeric ApMTAP

subunit including the hexahistidine tag (Figure 21A). This indicates that correctly folded ApMTAP was

expressed. The amount of ApMTAP, however, did not considerably depend on the choice of the

expression strain. On the other hand the cell density that was obtained with BL21 was significantly

higher than that obtained with the Rosetta strain (Figure 21B). Hence, only E. coli BL21 was used for

the following expression optimization.

Expression in EnPresso medium

The aim was to investigate whether the expression level of soluble ApMTAP can be further increased

by making use of EnPresso medium. Therefore the expression with E. coli BL21 at 30 and 37 °C with

three different levels of induction was analyzed. No strong effects of these parameters on the

amount of ApMTAP could be seen (Figure 22). However, the results show, that 20 µM IPTG is enough

for an efficient induction, and ApMTAP band intensity seemed to be higher in comparison to other

bands at 30 °C than at 37 °C.

(kDa) MBL21 Rosetta

S IN S IN

116

66

45

35

25

18

14

0

2

4

6

8

10

12

0 1 2 3 4 5 6

Op

tica

l den

sity

(6

00

nm

)

Cultivation time [h]

BL21

Rosetta

Induction

A) B)

Figure 21: ApMTAP expression in TB medium. (A) SDS-PAGE analysis of ApMTAP expressed with N-terminal hexahistidine tag (pKS2_ApMTAP) in E. coli BL21 or E. coli Rosetta and (B) growth curves of the cultivations. ApMTAP expression was induced with 20 µM IPTG at 37 °C 2.5 h after inoculation;cells were harvested 3 h after induction. The cell lysate was incubated at 85 °C prior to separation of soluble from insoluble protein fractions. The arrow indicates the bands corresponding to the molecular weight theoretically calculated for the monomeric ApMTAP subunit including hexahistidine tag (27 kDa); S = soluble, IN= insoluble protein fractions; M= molecular weight marker.

Another experiment was performed to check the influence of the time period of induction at 37 °C.

Therefore samples taken 6 h and 22 h after induction were analyzed. The results clearly show that

the expression level after the longer time period of induction, i.e. 22 h, lead to a higher expression

level of ApMTAP (Figure 23).

60 Recombinant expression of NPs

In contrast to DgPNP expression, where the amount of soluble protein could be significantly

increased in favour of insoluble protein with the help of EnPresso medium, the amount of soluble

ApMTAP per cell could not be increased with EnPresso medium in comparison to TB medium. On the

other hand, the volumetric yield of soluble ApMTAP could be considerably increased due to the

higher cell density reached: The optical density reached with EnPresso medium was more than two-

fold higher than that with TB medium (Figure 21B, Figure 23B).

(kDa) M

30 °C 37 °C

0 20 50 100 0 20 50 100

S S IN S IN S IN S S IN S IN S IN

116

66

45

35

25

18

14

IPTG (µM)

Figure 22: SDS-PAGE analysis of ApMTAP expression in EnPresso medium at 30 and 37 °C with 0 - 100 µM IPTG. E. coli BL21 pKS2_ApMTAP culture samples were harvested 25 h after induction; cell lysates were incubated at 85 °C prior to separation of soluble (S) from insoluble (IN) protein fractions. The arrow indicates the bands corresponding to the molecular weight theoretically calculated for the monomeric ApMTAP subunit including hexahistidine tag (27 kDa). M= molecular weight marker.

(kDa) M22 h 6 h

IN S IN S

116

66

45

35

25

18

14

0

5

10

15

20

25

30

0 10 20 30 40

Op

tica

l den

sity

(6

00

nm

)

Cultivation time [h]

Induction

A) B)

Figure 23: ApMTAP expression in EnPresso medium over time. (A) SDS-PAGE analysis of samples harvested 6 h and 22 h after induction and (B) corresponding growth curve. E. coli BL21 pKS2_ApMTAP was cultivated at 37 °C with 20 µM IPTG. The cell lysate was incubated at 85 °C prior to separation of soluble (S) from insoluble (IN) protein fractions. The arrow indicates bands corresponding to the molecular weight theoretically calculated for the monomeric ApMTAP subunit including hexahistidine tag (27 kDa); M= molecular weight marker.

Recombinant expression of NPs 61

Evidence for intersubunit disulfide bonds in ApMTAP

According to the theoretical predictions presented in Table 8, the cysteine in position 112 of ApMTAP

forms a disulfide bond. Therefore it seems likely that ApMTAP is stabilized by intersubunit disulfide

linkages. As it will be shown later (in section 4.2.1), based on amino acid sequence alignments

ApMTAP shows high similarity to MTAPI from S. solfataricus. The characterization of the

S. solfataricus enzyme in turn revealed a hexameric structures, in which three intersubunit disulfide

bonds link dimers to each other to form a hexamer (Appleby et al. 2001, Cacciapuoti et al. 1994). The

close evolutionary relationship of ApMTAP to MTAP from S. solfataricus hence corroborates the

presence of stabilizing intersubunit disulfide linkages, even though Cys129 instead of the predicted

Cys112 would be involved in the disulfide bond according to the amino acid sequence alignment.

Within the purification of ApMTAP, SDS-PAGE analysis revealed the presence of two bands. One

band corresponded to the molecular weight of the monomeric subunit (27 kDa), while the other

band represented a molecular weight of more than 116 kDa. It was confirmed by mass spectrometric

analysis (by Knut Büttner, Division of microbial physiology and molecular biology, Ernst-Moritz-Arndt-

University of Greifswald, Germany) that this second band also represents ApMTAP (data not shown).

For the reasons outlined in the previous paragraph we assumed that this second band represents

some oligomerization state of the monomeric subunit that is possibly stabilized by intersubunit

disulfide bonds. Further analysis indeed revealed that this second band was the predominant band

when ApMTAP samples were not treated with DTT prior to SDS-PAGE analysis. Conversely, the band

representing the molecular weight of the monomeric subunit was predominant after default sample

preparation that includes an incubation step at 95 °C in the presence of 200 µM DTT. Noteworthy

this phenomenon was not observed with the other proteins that were analyzed (Figure 24).

116

66

45

35

25

18

(kDa) M

- DTT treatment

TtP

yNP

Ap

MTA

P

DgP

NP

GtP

NP

GtP

yNP

TtP

yNP

Ap

MTA

P

DgP

NP

GtP

NP

GtP

yNP

Figure 24: Influence of DTT treatment on SDS-PAGE analysis of ApMTAP. Samples on the right side were prepared according to the standard protocol that includes DTT treatment. Samples on the left side were not treated with DTT. Arrows highlight the different positions of the bands representing ApMTAP in dependence of DTT treatment.

62 Recombinant expression of NPs

3.4.3. ApMTAP - summary and conclusions

Both the native ApMTAP sequence and an ApMTAP construct comprising a hexahistidine tag at the

N-terminus were cloned and overexpressed. Stable secondary 5′ mRNA structures impaired the

efficient expression of the native ApMTAP amino acid sequence initially. In analogy to DgPNP

expression optimization the yield of soluble ApMTAP was increased by a combination of 5′ mRNA

sequence optimization and high-temperature cultivation. The results further support the general

applicability of this strategy that is discussed in sections 3.3.2 and 3.3.4 in detail. Cloning ApMTAP

into a vector conferring N-terminally a hexahistidine tag (assigned here as pKS2_ApMTAP) proved to

be a good alternative to this approach for the purpose of this study: The ApMTAP gene product is still

thermostable, the expression is efficient and the hexahistidine tag enables the application of a simple

purification procedure that is favourable for following characterization studies (presented in

chapter 4 and 5).

Expression studies with pKS2_ApMTAP in E. coli BL21 or E. coli Rosetta as host yielded similar

amounts of recombinant protein, despite the presence of numerous rare codons. The volumetric

yield of ApMTAP expressed in BL21 was significantly higher in EnPresso medium than in TB medium,

due to the higher cell density reached. SDS-PAGE analysis and theoretical predictions strongly

support the presence of at least one intersubunit disulfide bond. However, the positive results

obtained here on the level of SDS-PAGE (and in the following sections on the basis of enzyme

activity) with E. coli BL21 or E. coli Rosetta suggest that it is not absolutely necessary to make use of

specialized expression systems. An example of such a specialized expression system is E. coli Origami,

a mutant strain with a mild oxidative cytoplasm that is offered by Novagen for the expression of

proteins containing disulfide bonds, but is generally more difficult to grow to high cell densities.

Noteworthy the successful cytoplasmic expression of ApMTAP in E. coli BL21 was recently also

reported by others (Zhu et al. 2012). The authors refer to a PNP, but comparison of the primers used

for the isolation reveals that the gene coding region coincides with the ApMTAP sequence here

under investigation.

The conclusions that can be drawn regarding the expression vector are basically the same as for

DgPNP expression. In both cases, expression of the target gene with hexahistidine tag in the pKS2

vector lead to presumably correctly folded protein in relatively high yield. Hence, the expression of

the following proteins will be directly started with this vector.

3.5. Expression of GtPNP

The gene coding for purine nucleoside phosphorylase (GenBank accession number EFG53380) was

isolated from G. thermoglucosidasius 11955. Since this microorganism shows optimal growth at 55 °C

(Suzuki et al. 1983), it was expected that the resulting protein is also temporarily stable at this

temperature. GtPNP was expressed in two different constructs. In the first vector, a hexahistidine tag

was fused to the C-terminus, in the second vector a hexahistidine tag was fused to the N-terminus.

Recombinant expression of NPs 63

3.5.1. GtPNP expression with C-terminal hexahistidine tag

Initially the GtPNP gene was cloned into the vector pET21a in such a way, that only the native gene is

expressed without any fusion partner. Expression resulted in high amount of both soluble GtPNP that

was retained in the soluble protein fraction after incubating the cell lysate at temperatures up to

65 °C (data not shown). To simplify the purification the vector was modified in such a way that a

hexahistidine tag is cloned to the C-terminus of the protein. Surprisingly, expression under similar

conditions as before (TB medium, 37 °C) resulted in almost exclusively insoluble recombinant GtPNP

(Figure 25).

3.5.2. GtPNP expression with N-terminal hexahistidine tag

Cloning and expression studies of DgPNP and ApMTAP have meanwhile shown that both proteins

were well expressed from the vector pKS2 that confers an N-terminal hexahistidine tag. The same

vector was used now for GtPNP expression. E. coli BL21 was transformed with pKS2_GtPNP,

cultivated at 37 °C in TB medium and induced with 100 µM IPTG. Under these conditions, GtPNP was

expressed in very high yield in predominantly soluble form. In addition the gene product could be still

found in the soluble protein fraction after incubation at 65 °C for 10 min. These positive results made

any further expression optimization unnecessary.

(kDa) M S IN

116

66

45

35

25

18

14

Induction

0

2

4

6

8

10

12

14

0 2 4 6

Op

tica

l den

sity

(6

00

nm

)

Cultivation time [h]

A) B)

Figure 25: Expression of GtPNP with C-terminal hexahistidine tag. (A) SDS-PAGE analysis of E. coli Rosetta2 (DE3) pET21a_GtPNP culture sample (B) growth curve of the cultivation. Cells cultivated in TB medium at 37 °C; induction of GtPNP expression with 100 µM IPTG; harvest 3 h after induction. The arrow indicates the bands corresponding to the molecular weight theoretically calculated for the monomeric GtPNP subunit including hexahistidine tag (27.6 kDa). S= soluble, IN= insoluble protein fraction; M= molecular weight marker.

3.5.3. GtPNP expression - summary and conclusion

The expression studies on GtPNP have further highlighted the critical role of fusion tags on the

functional expression of nucleoside phosphorylases: Apparently not only the length (see

section3.3.1), but also the position of the fusion tag is decisive for proper folding. While the C-

64 Recombinant expression of NPs

terminal hexahistidine tag lead to aggregation, GtPNP was very efficiently expressed in soluble form

from pKS2 with N-terminal fusion of the hexahistidine tag. The high yield of functional GtPNP made

further expression optimization superfluous.

(kDa) M S IN

116

66

45

35

25

18

14

0

4

8

12

16

0 2 4 6 8O

pti

cal d

ensi

ty (

60

0 n

m)

Cultivation time [h]

Induction

A) B)

Figure 26: Expression of GtPNP with N-terminal hexahistidine tag. (A) SDS-PAGE analysis of E. coli BL21 pKS2_GtPNP samples (B) growth curve of the cultivation. GtPNP expression in TB medium was induced with 100 µM IPTG at 37 °C. Samples were harvested 3.5 h after induction; the cell lysate was incubated at 65 °C prior to separation of soluble (S) from insoluble (IN) protein fractions. The arrow indicates the bands corresponding to the molecular weight theoretically calculated for the monomeric GtPNP subunit including hexahistidine tag (27.6 kDa); M= molecular weight marker.

3.6. Expression of GtPyNP

Next to the purine nucleoside phosphorylase the pyrimidine nucleoside phosphorylase of

G. thermoglucosidasius 11955 was studied. It was likewise expected that the recombinant protein is

stable at least at 55 °C. GtPyNP was expressed with and without N-terminal hexahistidine tag.

Parts of the expression study on GtPyNP have been previously published (Szeker et al. 2012).

3.6.1. Chemical lysis buffer decreases apparent thermal stability of GtPyNP

In analogy to GtPNP, GtPyNP was first expressed in the vector pET21a in such a way that no tag is

fused to the protein. Expression in TB medium under standard conditions (37 °C, 100 µM IPTG)

resulted in a high yield of recombinant GtPyNP that was almost exclusively produced in soluble form

(Figure 27). However, unexpectedly and in contrast to GtPNP, the gene product appeared to be not

thermostable. When the cell lysate was incubated at temperatures exceeding 45 °C, GtPyNP was not

completely retained in the soluble protein fraction.

With the assumption that GtPyNP did not obtain its native conformation and hence lacked

thermostability, two measures were undertaken to overcome this problem. Firstly, the same plasmid

(pET21a_GtPyNP) was expressed in E. coli Origami cells. The idea behind was that disulfide bond

formation could have been necessary to obtain thermostable GtPyNP. In contrast to usual E. coli

Recombinant expression of NPs 65

expression strains, Origami promotes disulfide bond formation of recombinant proteins expressed in

the cytoplasm of E. coli. The second measure undertaken applied to the amino acid sequence.

Compared to the data bank sequence, the gene cloned here encoded glutamine (side chain not

charged at physiological pH) on position 214 instead of arginine (positively charged side chain). This

mutation could have been the result of errors that occurred during the cloning procedure, for

example during the amplification of the gene via PCR. Hence, the coding region was modified by site-

directed mutagenesis, to match the respective databank entry (gene bank accession number

ZP_06809030). However, both measures did not change the apparent lack of thermostability as

shown in Figure 27A.

Finally the solution for the problem was found in the nature of the buffer, in which the cell lysate was

incubated to estimate the stability at high temperatures. So far, cells were disrupted through

treatment with a chemical lysis buffer, namely Bugbuster® protein extraction reagent (Novagen).

However, in following experiments cell pellets were resuspended in a phosphate buffer (50 mM,

pH 8) and disrupted by sonication. Investigation of cell lysates obtained with this method showed

that GtPyNP appeared to be stable at temperatures up to 60 °C now (Figure 27B).

3.6.2. GtPyNP expression with N-terminal hexahistidine tag

The positive results obtained with the pKS2 vector prompted us to use the same vector for GtPyNP

expression. E. coli BL21 pKS2_GtPyNP was cultured under standard conditions (TB medium, 37 °C)

and recombinant protein expression was induced by the addition of 100 µM IPTG. GtPyNP was

expressed in very high yield in predominantly soluble form and was retained in the soluble protein

fraction after incubation of the cell lysate at 60 °C. Further optimization was considered as

unnecessary.

(kDa) M- 45 °C 50 °C 55 °C

S IN S IN S IN S IN

116

66

45

35

25

18

14

(kDa) M55 °C 60 °C 65 °C 70 °C 75 °C

S IN S IN S IN S IN S IN

116

66

45

35

25

18

14

A) B)

Figure 27: SDS-PAGE analysis of thermo-treated E. coli BL21 pKS2_GtPyNP culture samples. Cell disruption was performed by chemical means with lysis buffer (A) or by sonication of cells resuspended in phosphate buffer (B). Soluble (S) and insoluble (IN) fractions of culture samples that were treated at the indicated temperatures for 10 min are shown. Arrows indicate the bands corresponding to the molecular weight theoretically calculated for the monomeric GtPyNP subunit (46 kDa); M= molecular weight marke.

66 Recombinant expression of NPs

A) (kDa) M S IN

116

66

45

35

25

18

14

0

2

4

6

8

10

12

0 2 4 6 8

Op

tica

l den

sity

(6

00

nm

)

Cultivation time [h]

B)

Induction

Figure 28: Expression of GtPyNP with N-terminal hexahistidine tag. (A) SDS-PAGE analysis of E. coli BL21 pKS2_GtPyNP culture sample, (B) growth curve of the cultivation. GtPyNP expression in TB medium was induced with 100 µM IPTG at 37 °C. Samples were harvested 3.5 h after induction. The cell lysate was incubated at 60 °C prior to the separation of the soluble (S) from the insoluble (IN) protein fraction. The arrow indicates the bands corresponding to the molecular weight theoretically calculated for the monomeric GtPyNP subunit including hexahistidine tag (47.6 kDa). M= molecular weight marker

3.6.3. GtPyNP expression - summary and conclusions

So far, incubation of the cell lysates of expression cultures at high temperatures and subsequent

SDS-PAGE analysis has served as quick tool to evaluate whether the target protein is properly folded.

However, this section has demonstrated the limitations of this procedure, since it was shown that

the buffer composition dramatically influenced the apparent stability of GtPyNP. In chapter 4- that

deals with the characterization of the target enzymes, stability issues will be investigated in more

detail.

As observed for all other target proteins investigated so far, the pKS2 vector was found to be a good

expression vector for GtPyNP, too. Indeed, the expression efficiency under standard conditions was

so high, that optimization was considered as unnecessary and investigations on GtPyNP could

proceed directly with the characterization of the recombinant protein.

3.7. Expression of TtPyNP

The gene coding for pyrimidine nucleoside phosphorylase was isolated from

T. thermophilus HB27.Since this microorganism shows optimal growth at around 68 °C (Oshima and

Imahori 1974), it was expected that the resulting recombinant protein (TtPyNP) is also temporarily

stable at this temperature.

Parts of the results presented in this section have been previously published (Szeker et al. 2012).

Recombinant expression of NPs 67

3.7.1. TtPyNP expression with N-terminal hexahistidine tag

TtPyNP was cloned in vector pKS2. The resulting expression plasmid (pKS2_TtPyNP) was transformed

in E. coli BL21 and E. coli Rosetta. Initial studies showed that TtPyNP was predominantly expressed in

soluble form and was completely retained in the soluble fraction after incubation at 80 °C (data not

shown).

The expression level of TtPyNP was generally poor. In order to optimize the yield of soluble TtPyNP

different cultivation media were studied. Experiments within the course of this study revealed that

the most reproducible result was obtained through the use of media with enzyme based glucose

delivery (EnBase Flo, EnPresso) in contrast to conventional complex media (LB, TB) (data not shown).

In Figure 29 an exemplary expression study in the different media is shown in which the best result

was obtained with EnPresso medium. The figure also indicates that despite the abundance of rare

codons in the TtPyNP gene, the expression level per cell could not be significantly increased by the

use of the E. coli Rosetta strain in which rare tRNAs are overexpressed.

(kDa) MRosetta2 BL21

LB TB EP LB TB EP

116

66

45

35

25

18

Figure 29: SDS-PAGE analysis of pKS2_TtPyNP expression in E. coli Rosetta2 or BL21 in different media. The cultivations were performed at 37 °C in LB, TB, or EnPresso (EP) medium; induction of TtPyNP expression with 1 mM IPTG. Soluble fractions of samples were treated for 5 min at 80 °C after cell disruption. The arrow indicates the bands corresponding to the molecular weight theoretically calculated for the monomeric TtPyNP subunit including hexahistidine tag (approx. 47 kDa); M= molecular weight marke.

Further investigation on TtPyNP expression in E. coli BL21 cultivated in EnBase Flo medium showed,

that the expression level increased only moderately with the expression time. Nevertheless, a higher

volumetric yield could be obtained by extending the time period after induction (24 h), due to the

higher optical density reached: The final OD600 obtained with EnBase Flo was 24 (Figure 30); instead,

with TB medium only OD600 =11 was reached (data not shown).

68 Recombinant expression of NPs

0

4

8

12

16

20

24

0 10 20 30 40

Op

tica

l den

sity

(6

00

nm

)

Cultivation time [h]

(kDa) M

S IN

Thermo -treatment

-Thermo-

treatment-

0 8 24 0 8 24 0 8 24 0 8 24

116

66

45

35

25

18

14

A) B)

Induction

Figure 30: Expression pKS2_TtPyNP E. coli BL21 in dependence of expression time. (A) SDS-PAGE analysis, (B) growth curve of the cultivation. The cultivation was performed at 30 °C in EnBase Flo medium, induction with 1 mM IPTG. Cells were harvested 0, 8, and 24 h after induction. Soluble (S) and insoluble (IN) protein fractions were analyzed before and after thermo-treatment for 5 min at 80 °C. The arrow indicates the bands corresponding to the molecular weight theoretically calculated for the monomeric TtPyNP subunit including hexahistidine tag (47 kDa); M= molecular weight marke.

3.7.2. TtPyNP expression - summary and conclusions

TtPyNP was successfully expressed with N-terminal hexahistidine tag. The gene product was retained

in the soluble protein fraction after heat-treatment at 80 °C, which indicates that TtPyNP was

correctly folded. The expression level was generally poor. Expression optimization resulted in slightly

higher expression levels, but the final yield of TtPyNP was significantly lower compared to that of the

other proteins investigated in this study, for example GtPyNP. Nevertheless, suitable expression

conditions were found for production of a sufficient amount of TtPyNP for characterization studies

and synthetic reactions as described in chapters 4 and 5.

3.8. Expression of ApUP

The gene coding for uridine phosphorylase was the second gene isolated from the hyperthermophilic

archaeon A. pernix in this study. Due to its origin, the resulting recombinant protein (ApUP) was

hence expected to be also temporarily stable at 90 - 95 °C, where A. pernix shows optimal growth.

The ApUP sequence is characterized by a number of peculiarities that might pose a challenge for the

successful recombinant expression. According to the theoretical predictions listed in Table 8, the

ApUP gene sequence contains numerous rare codons. Moreover, 2 cysteine residues per subunit

were predicted to be in the bonding state, which suggests that inter- or intrasubunit disulfide bonds

play a role for stabilization. One of the cysteines predicted to be in the bonding state (Cys225) is part

of a CXC motif that was reported to be required for folding in the disulfide bond containing proteins

PfPNP (PNP from P. furiosus) and SsMTAPII (MTAPII from S. solfataricus) (Cacciapuoti et al. 2009).

Furthermore, the gene sequence contains two tandemly repeated AGG triplets. This AGGAGG

sequence motif is part of the natural consensus Shine-Dalgarno sequence and was reported to inhibit

Recombinant expression of NPs 69

recombinant protein expression by competing with the “real” Shine-Dalgarno sequence for ribosome

binding (Ivanov et al. 1992, Jin et al. 2006). Noteworthy the gene coding sequence, as it was

annotated originally started with the initiating codon located downstream of this internal ribosomal

binding site. Just after the re-annotation of the A. pernix genome in 2006 the full length ApUP gene

starting before the AGGAGG motif was assigned (Yamazaki et al. 2006). Finally, when the native gene

is cloned in the expression vector used here (resulting in pKS1_ApUP) the free energy of the 5′ mRNA

folding is similarly low (-17.0 kcal mol-1) as reported before for ApMTAP (Table 10), suggesting that

secondary structures might impair the efficient expression.

3.8.1. Expression of ApUP without tag

For the first series of experiments the native ApUP gene (without fusion tag) was overexpressed in

three different E. coli expression strains. These strains include E. coli BL21 as standard strain, E. coli

Origami – that allows disulfide bond formation in the cytoplasm and E. coli Rosetta-gami in which

additionally rare tRNAs are overexpressed to compensate for the abundance of codons in the ApUP

gene that are only rarely used in E. coli.

No thermo-treatment Thermo-treatment at 90 °C

Origami Rosetta-gami BL21 Origami Rosetta-gami BL21

+P +P +P +P +P +P

(kDa) M S IN S IN S IN S IN S IN S IN S IN S IN S IN S IN S IN S IN

116

66

45

35

25

18

14

Figure 31: SDS-PAGE analysis of ApUP expression in E. coli Origami, Rosetta-gami, and BL21. Expression strains harbouring pKS1_ApUP (+P) and empty strains without expression plasmid were cultivated at 37 °C in TB medium, induction of ApUP expression with 20 µM IPTG. Soluble (S) and insoluble (IN) protein fractions were analyzed before (left side) and after thermo-treatment at 90 °C (right side). The arrow indicates the bands corresponding to the molecular weight theoretically calculated for the monomeric ApUP subunit (30.3 kDa); M= molecular weight marker

Expression experiments were performed in TB medium at 37 °C, Protein expression was induced by

adding IPTG to a final concentration of 20 µM. The same expression strains but without recombinant

plasmids were cultivated in parallel as a reference. Soluble and insoluble protein fractions were

analyzed by SDS-PAGE before and after heat treatment at 90 °C (Figure 31). A protein with a

molecular size according to the theoretically calculated value for monomeric ApUP was found in the

soluble protein fraction before and after the heat treatment. It was not totally clear, whether this

protein really represents ApUP since a thermostable protein with similar size was also found in

70 Recombinant expression of NPs

“empty” Origami and Rosetta-gami strains. In BL21, however, the expression was more pronounced;

a clear band was observed in BL21 with the recombinant plasmid exclusively. On the other hand,

SDS-PAGE analysis also indicates that the gene product is not totally stable at 90 °C, since the amount

of soluble ApUP seems to decrease after the thermo-treatment. It remains to be interesting whether

the reason can be found in the conditions of the thermo-treatment that might be too harsh (too high

temperature, improper buffer), or in the recombinant protein itself that might have not obtained its

native conformation.

ApUP ApUP1

No thermo-treatment Thermo-treatment No thermo-treatment Thermo-treatment

30 °C 37 °C 42 °C 30 °C 37 °C 42 °C 30 °C 37 °C 42 °C 30 °C 37 °C 42 °C

(kDa) M S IN S IN S IN S IN S IN S IN S IN S IN S IN S IN S IN S IN

116

66

45

35

25

18

14

Figure 32: SDS-PAGE analysis of ApUP expression in E. coli BL21 with the original gene sequence (ApUP) and the optimized sequence, lacking an internal ribosomal binding site (ApUP1). Cultivations were performed at 30, 37, and 42 °C in TB medium; induction with 20 µM IPTG. Soluble (S) and insoluble (IN) protein fractions were analyzed before and after thermo-treatment which was initially performed at 85 °C (ApUP) or 90 °C (ApUP1). The arrow indicates the bands corresponding to the molecular weight theoretically calculated for the monomeric ApUP subunit (30.3 kDa); M= molecular weight marker

A closer look at the ApUP gene revealed the presence of a sequence motif that could serve as

internal ribosomal binding site. In this case, a second, smaller protein would be produced from the

same mRNA molecule. It is possible that in the expression experiment with BL21 and pKS1_ApUP

shown in Figure 31 a mixture of the thermostable full length protein and a slightly shorter,

thermolabile fragment was produced. This would also be an explanation for the apparent loss of

soluble ApUP after thermo-treatment. In order to investigate this hypothesis, the potential internal

ribosomal binding site was removed by silent mutations (AGGAGG CGGCGT); the resulting

expression vector was assigned as pKS1_ApUP1. In addition, sequence analysis of pKS1_ApUP mRNA

has also revealed, that stable secondary structures (free energy of formation of -17 kcal mol-1 at

37 °C) are likely to occur in the 5′ mRNA region. The following expression experiments were hence

performed at different expression temperatures, to investigate whether the yield of recombinant

protein can be increased by high-temperature cultivation, as done before (sections 3.3.2 and 3.4.1).

SDS-PAGE analysis of the recombinant cultures (Figure 32) indeed showed that the yield of soluble

Recombinant expression of NPs 71

ApUP was slightly better at 42 °C. However, the removal of the potential internal ribosomal binding

site did not have a visible positive effect on ApUP expression.

3.8.2. Expression of ApUP with N-terminal hexahistidine tag

In order to produce high amounts of ApUP that can be easily purified for characterization studies,

ApUP1 was also expressed with an N-terminal hexahistidine tag (pKS2_ApUP1). Next to the full

length protein, also the shorter fragment starting after the internal ribosomal binding site was cloned

in this way, resulting in a vector assigned as pKS2_ApUPsh. The respective open reading frame that

starts after the internal ribosomal binding site and coincides with the coding region as it was

annotated for the ApUP gene in the past. Since both the older and the topical annotation are based

on theoretical predictions only, both open reading frames were considered for the final investigation

on ApUP expression of this study. The expression conditions were chosen as before for the native

gene. However, SDS-PAGE analysis of the culture samples revealed that both ApUP constructs with

the N-terminal hexahistidine tag resulted exclusively in the formation of insoluble protein,

independent of the length of the coding region cloned (ApUP1, ApUPsh) and the cultivation

temperature.

30 °C 37 °C 42 °C

ApUP1 ApUPsh ApUP1 ApUPsh ApUP1 ApUPsh

(kDa) M T S T S T S T S T S T S

116

66

45

35

25

18

14

Figure 33: SDS-PAGE analysis of ApUP expression with N-terminal hexahistidine tag in E. coli Rosetta. Both the expression of the full length protein (ApUP1) and the shorter fragment that represents the former ApUP annotation (ApUPsh) was investigated. Soluble (S) and total (T) protein fractions were analyzed after thermo-treatment at 90 °C. The arrows indicate the vertical position of bands presumably representing the full length protein and the shorter fragment.M= molecular weight marker

3.8.3. ApUP expression - summary and discussion

ApUP expression proved to be very challenging. Even at very low expression levels, a high amount of

recombinant protein was expressed in insoluble form. Various aspects were considered for

expression optimization: Different expression strains were tested, an apparent internal ribosomal

binding site was removed and finally also a shorter version of the ApUP gene that represents the old

72 Recombinant expression of NPs

ApUP annotation was considered. The best result, i.e. the highest amount of soluble and

thermostable ApUP, was obtained by the expression of the full-length ApUP gene without any tag.

Fusion of an N-terminal hexahistidine tag - that was a successful strategy to obtain high yields of all

the other proteins studied here – resulted in the synthesis of ApUP in exclusively insoluble form. It

might be of interest for future experiments to study i) whether it is possible to express ApUP with N-

terminal hexahistidine tag by further reducing the expression level and making use of media with

enzyme controlled glucose delivery, ii) whether other lysis buffers would lead to a different result

(having in mind the experiences with GtPyNP expression analysis) and iii) to establish another

purification method and purify ApUP expressed without hexahistidine tag. However, since in the

meantime ApUP was successfully recombinantly expressed by others (Montilla Arevalo et al. 2011)

and investigated with respect to its application as biocatalyst for the synthesis of nucleosides, the

focus of this study shifted more to the other 5 target enzymes, already successfully expressed, and

ApUP expression was not further investigated. It is interesting to note, that the authors of the

invention claim to have successfully expressed uridine phosphorylase from A. pernix by making use of

an expression vector in which thioredoxin is fused to the N-terminal end and a hexahistidine tag to

the C-terminal end, and that the ApUP gene sequence that was cloned represents the older

annotation of the gene, that is the shorter fragment starting after the internal ribosomal binding site.

In another very recent study ApUP expression from plasmid pET30a in E. coli BL21 was reported.

According to the primer sequences used, here the full length gene was cloned (Zhu et al. 2012).

Apparently a hexahistidine tag was fused to the C-terminus, since the reverse primer lacks a stop

codon and a hexahistidine tag is indicated downstream of the restriction site used in pET30a.

3.9. Recombinant expression of NPs - summary and conclusions

Six nucleoside phosphorylases were selected from four different thermophilic microorganisms that

show optimal growth between 50 °C and 95 °C. With the aim in view to generate biocatalysts for

transglycosylation of nucleosides from pyrimidine nucleoside donor to purine base acceptor, the

selection comprises enzymes with specificity towards purine and pyrimidine nucleosides. Five of the

six target proteins were successfully overexpressed in E. coli in moderate or high yield. Based on

thermostability studies with SDS-PAGE analysis these enzymes obtained the correctly folded form.

Overexpression of ApUP proved to be challenging. A variety of attempts to produce thermostable

ApUP failed and finally investigations on expression optimization was discontinued for the sake of

time and owing to the fact that the successful expression and biocatalytic application of ApUP was

already reported in the meantime by others (Montilla Arevalo et al. 2011, Zhu et al. 2012). A number

of obstacles were encountered and had to be overcome also for the five proteins that were

successfully overexpressed. The initially selected expression system proved to be not suitable, most

likely because too many additional amino acids were conferred to the N-terminus of the target

sequence, which possibly precluded the correct folding of the monomeric subunits or impaired the

oligomerization step, respectively. After the elimination of additional N-terminal amino acids, the

formation of stable secondary 5′ mRNA structures impaired the efficient expression of wild type

sequences without N-terminal fusion tag. This hurdle was tackled by an approach involving sequence

Recombinant expression of NPs 73

optimization and high-temperature cultivation. This strategy might also be useful for the expression

of other native gene sequences without N-terminal fusion, especially when derived from

thermophilic microorganisms. The choice of an expression vector that confers N-terminal a

hexahistidine tag proved to be a good alternative and was eventually used for all the five target

proteins. For the expression of a single protein that was not hampered by the formation of stable

secondary 5′ mRNA (GtPNP), also the C-terminal fusion of a hexahistidine tag was investigated, but

this strategy lead to aggregation of the protein product. One of the candidates (GtPyNP) was readily

expressed in soluble form but appeared to lack thermal stability. Closer investigations revealed that

the chemical lysis buffer used had an adverse effect on the thermostability of this protein.

Remarkably an adverse effect of the same lysis buffer was not observed for any other target protein.

Some proteins were readily expressed in soluble form and in very high yield (GtPNP, GtPyNP) while

the expression levels of other proteins was poor (TtPyNP), leading to a low volumetric yield or the

ratio of soluble (presumably correctly folded) to insoluble (presumably not correctly folded) protein

was unfavourable (DgPNP, ApMTAP) under standard conditions (TB medium, 37 °C, 100 µM IPTG).

Both the volumetric yield and the ratio of soluble to insoluble protein could be improved by fine-

tuning the IPTG concentration and by making use of media in which the growth rate is restricted by

enzyme based glucose delivery. Despite the abundance of rare codons in some of the target

sequences, the use of an E. coli strain overproducing rare tRNAs (Rosetta) had either no positive

effect or yielded only moderate improvement that was offset by the decreased final cell density

reached in comparison to the standard expression strain used (BL21).

In the final expression system all target proteins are hence expressed in E. coli BL21 from the same

vector backbone that confers an N-terminal hexahistidine tag. This strategy also paves the way for a

simple purification strategy for all target proteins, a prerequisite for rapidly progressing to the next

work package, the characterization of the biocatalytical properties of the purified enzymes.

4. Characterization of thermostable NPs

The final aim of this study was the synthesis of modified nucleosides by an enzymatic

transglycosylation reaction employing thermostable nucleoside phosphorylases. In order to transfer

a pentofuranosyl moiety from pyrimidine nucleoside to purine base two types of enzymes are

required. In a first step an enzyme with PyNP activity phosphorolytically cleaves the donor

nucleoside; in the second step an enzyme with PNP activity catalyzes the glycosyl bond formation

between the intermediate product (pentofuranosyl-1-phosphate) and the purine base.

This chapter is devoted to the characterization of thermostable NPs that could be employed in this

reaction scheme described. In focus are thermal properties (temperature optimum, stability) and

substrate specificities towards natural and artificial nucleosides. The results help to evaluate which

enzymes are most suitable for a specific catalytic reaction. Moreover, the data help to define the

process operating windows for transglycosylations employing the studied thermostable biocatalysts.

4.1. Thermostable PyNPs

PyNPs have been hardly studied in detail. Only few examples can be found in the scientific literature:

PyNP from B. subtilis (Gao et al. 2006), G. stearothermophilus (Hamamoto et al. 1996, Hori et al.

1990, Saunders et al. 1969) and from T. thermophilus (Shimizu and Kunishima 2007). With respect to

practical use as biocatalyst, only the G. stearothermophilus enzyme has been described (Taran et al.

2009).

The following sections deal with the characterization of recombinantly expressed PyNPs derived from

the thermophilic microorganisms G. thermoglucosidasius 11955 (GtPyNP) and T. thermophilus HB27

(TtPyNP). Albeit the expression and crystallization of PyNP from T. thermophilus HB8 has been

reported (Shimizu and Kunishima 2007), information about the biocatalytic characterization is not

available. Here, data reporting on thermal properties and steady state kinetics for the natural

substrates uridine and thymidine is presented. Of special interest is furthermore the ability of the

thermostable PyNPs to phosphorolyze unnatural pyrimidine nucleosides. Specifically the two 2′-

fluorosubstituted pyrimidine nucleosides 2′-deoxy-2′-fluorouridine (dUrd2′F) and 1-(2-deoxy-2-fluoro-

-D-arabinofuranosyl)uracil (dUrd2′F) are tested as substrates. Both compounds are highly interesting

pentofuranosyl donors in enzymatic transglycosylation reactions aiming at the synthesis of the

respective sugar-modified purine nucleosides. A precondition is however, the availability of PyNPs

that can efficiently catalyze the phosphorolysis of these donor nucleosides.

Major parts of the results presented in this chapter, have been published previously (Szeker et al.

2012). Thermal unfolding of TtPyNP (Figure 37) was studied by Thomas Schwab from the laboratory

of Reinhard Sterner (Institute of Biophysics and Physical Biochemistry, University of Regensburg,

Germany).

76 Characterization of thermostable NPs

4.1.1. Homology modelling

Three entries of solved crystal structures of PyNPs can be found in the Protein Database Bank

(http://www.rcsb.org/pdb/) (Bernstein et al. 1977). Both target enzyme sequences were blasted

against the amino acid sequences belonging to these PDB entries. GtPyNP aligned best with the PyNP

from G. stearothermophilus ATCC 12980 (assigned here as GsPyNP, PDB ID: 1BRW) with 78 %

sequence identity. TtPyNP aligned best with the PyNP from T. thermophilus HB8 (PDB ID: 2DSJ), to

which it is almost identical (approx. 98 % sequence identity) but showed also a high degree of

sequence identity (approx. 50 %) to GsPyNP. The amino acid sequence alignments of GsPyNP,

GtPyNP, TtPyNP, and E. coli thymidine phosphorylase (EcTP) are shown in Figure 34. Amino acids that

have been described to be involved in substrate binding or in the catalytic mechanism, respectively,

are indicated (Mendieta et al. 2004, Pugmire and Ealick 1998).

The chain B of GsPyNP (PDB ID: 1BRW) was used as template to model both TtPyNP and GtPyNP. The

structure of GsPyNP was elucidated in its closed conformation (Pugmire and Ealick 1998), uracil and

the phosphate ion can be seen in the active site pocket. The structural folds of the models of GtPyNP

and TtPyNP are almost identical to the template structure GsPyNP, which can be seen in the

superposition of the secondary structure elements Figure 35.

4.1.2. Thermal characteristics

In order to assess the optimal temperature range at which the PyNPs can be used in potential bio-

synthetic applications, the optimal temperatures and thermal stabilities of the enzymes were

studied.

GtPyNP showed a temperature optimum of 60 °C, while the relative activity of TtPyNP increased with

the reaction temperature up to the highest temperature tested (99 °C, Figure 36A). An apparent

melting temperature of ≥ 102 °C and 103 °C was determined by circular dichroism and differential

scanning calorimetry, respectively (Figure 37). Hence, we assume that the temperature optimum of

TtPyNP is in the range of 95 °C – 103 °C.

The stability half life of GtPyNP was determined to be 1.6 h at 70 °C; while at 60 °C no significant loss

of activity could be seen within 16 h of incubation. The stability half life of TtPyNP exceeds 23 h at

80 °C (Figure 36B). At 90 °C TtPyNP was almost completely deactivated within 6 hours (data not

shown).

Compared to other reported enzymes with pyrimidine nucleoside phosphorylase activity, including

UPs and TPs, the thermal stability and the temperature optimum of TtPyNP is extremely high (Table

11). It seems therefore appealing to further investigate whether this highly thermostable biocatalyst

can be expediently used for enzymatic transglycosylations reactions aiming at the synthesis of

modified nucleosides. Based on the results obtained here, further characterization of GtPyNP and

TtPyNP will be performed at 60 °C and 80 °C, respectively. These temperatures seem adequate since

both enzymes show sufficiently high activity and at the same time stay stable for prolonged

incubation times.

Characterization of thermostable NPs 77

Figure 34: Multiple sequence alignment of PyNP from G. stearothermophilus (GsPyNP), G. thermoglucosidasius (GtPyNP), T. thermophilus (TtPyNP), and TP from E. coli (EcTP). Shading represents the degree of sequence identity. Residues of the active site pocket are highlighted. Figure taken from (Szeker et al. 2012).

Figure 35: Three-dimensional structure models. The superposition of GtPyNP (purple), TtPyNP (yellow), and GsPyNP (grey) 3D structures are shown. Structural models of GtPyNP and TtPyNP were built based on homology modelling using the structure of GsPyNP as template. Figure taken from (Szeker et al. 2012).

78 Characterization of thermostable NPs

Figure 36: Thermal characteristics. Relative activity of GtPyNP and TtPyNP over the reaction temperature (A) where the highest reaction rate determined was set to 100 % for each enzyme. Thermostability (B) was investigated by incubating protein samples for defined time intervals and subsequently determining the residual activity, where the activity of enzyme samples that were not thermo-treated was set to 100 %. Figure adapted from (Szeker et al. 2012).

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

20 30 40 50 60 70 80 90 100 110

No

rmal

ized

elli

pti

city

Temperature [°C]

-40

-30

-20

-10

0

10

20

30

40

75 80 85 90 95 100 105 110 115 120

c p[k

cal K

-1m

ol-1

]

Temperature [°C]

A) B)

Figure 37: Thermal unfolding of TtPyNP. Thermal unfolding trace monitored by loss of the far-UV circular dichroism signal at 220 nm provided an apparent melting temperature of at least 102 °C (A). The apparent melting temperature determined by differential scanning calorimetry (B) was 103 °C. Figure adapted from (Szeker et al. 2012).

4.1.3. Kinetic parameters

In order to compare the thermostable PyNPs prepared in this study with PyNPs described by others,

parameters describing substrate specificity and catalytic efficiency were studied (Table 12).

Michaelis-Menten kinetics for the natural substrates uridine and thymidine were investigated by

determining the reaction rates at different substrate concentration. The amount of phosphate, that

Characterization of thermostable NPs 79

in fact represents the second substrate, was kept constant whereby the concentration (50 mM)

considerably exceeded the concentration of the nucleoside substrate. Enzyme reactions with GtPyNP

were performed at 60 °C and with TtPyNP at 80 °C. The previous section has shown that both

enzymes are stable over extended time periods at these temperatures.

Table 11. Thermal characteristics of reported PyNP, UP, TP. Table adapted from (Szeker et al. 2012).

Enzyme Organism Thermal stability (t1/2) Temp. optimum Reference

TP E. coli <10 min (55 °C) (Krenitsky and Tuttle

1982)

UP E. coli 9.9 h (60 °C) 40 °C (Visser et al. 2011)

PyNP G. thermoglucosidasius 1.6 h (70 °C) 60 °C This study

UP* E. coli 3.3 h (70 °C) 60 °C (Visser et al. 2011)

UP Enterobacter aerogenes 1 week (60 °C) 65 °C (Utagawa et al. 1985a)

PyNP G. stearothermophilus 25 min (70 °C) 70 °C (Hamamoto et al. 1996,

Hori et al. 1990)

UP Erwinia carotovora - 70 °C (Shirae and Yokozeki

1991)

PyNP T. thermophilus >23 h (80 °C) > 95 °C This study

* UP from E. coli was engineered for enhanced stability, for details see (Visser et al. 2011)

The Michaelis-Menten constant (Km) is defined as the substrate concentration at which the reaction

rate is half of the maximal velocity. It is a common parameter used to describe the affinity of an

enzyme towards a specific substrate. The estimated Km values of TtPyNP for the natural substrates

uridine and thymidine are slightly lower than the Km values determined by Hori and co-workers(Hori

et al. 1990) for GsPyNP. In contrast to TtPyNP, GtPyNP is characterized by extremely low substrate

affinities (high Km values) towards both natural substrates.

The catalytic efficiency of an enzyme can be described by the ratio of kcat/Km (Copeland 2000). These

ratios are 15 times (substrate uridine) and 25 times (substrate thymidine) higher for TtPyNP than for

GtPyNP. The kcat/Km ratio is also a measure to compare an enzyme’s specificity towards different

substrates (Copeland 2000). The results of the present study show that both enzymes are more

specific for uridine than for thymidine, but the difference in specificity is more pronounced for

GtPyNP: the kcat/Km ratio is 2fold higher for uridine than for thymidine. In contrast, the kcat/Km ratio of

TtPyNP is only 1.23 fold higher for uridine vs that of thymidine.

In applications, where high substrate concentrations are used (cs>> Km) the kcat value alone, also

referred to as turnover number, may be the most appropriate parameter describing the efficiency of

the biocatalyst. It describes on how many substrate molecules a single enzyme molecule is acting in

80 Characterization of thermostable NPs

one second. The turnover numbers (kcat) of GtPyNP and TtPyNP are in similar range, with uridine as

substrate. By contrast, the turnover numbers for thymidine differ significantly, in favour of thymidine

phosphorolysis by TtPyNP. The kcat value of TtPyNP for thymidine is also unusually high in comparison

to alternative enzymes that are used for the phosphorolysis of pyrimidine nucleosides, e.g. EcUP and

EcTP (Table 12).

Table 12. Kinetic parameters of PyNPs, EcTP and EcUP. Table adapted from (Szeker et al. 2012).

The estimation errors of the Km and kcat values determined in this study were not higher than 10 %. Kinetic parameters were determined at 60 °C pH 7.0 (GtPyNP, GsPyNP), at 80 °C pH7.0 (TtPyNP),at 25 °C pH 6.5 (EcTP), and at 25 °C pH 7.5 (EcUP). (-) Data not indicated.

4.1.4. Phosphorolysis of 2′-fluorosubstituted pyrimidine nucleosides

Of particular interest is the potential of both thermostable PyNPs as biocatalysts in the synthesis of

modified nucleosides. With this aim in view, the phosphorolysis of natural pyrimidine nucleoside

substrates (thymidine and uridine) and their sugar modified analogues, i.e. dUrd2′F and dUrd2′F, were

investigated. These substrates can be used as pentofuranosyl donors in enzymatic

transglycosylations aiming at the synthesis of pharmaceutically valuable 2′-fluorosubstituted purine

nucleosides. With this strategy dUrd2′F served as a substrate for the enzymatic synthesis of 2′-deoxy-

2′-fluoroguanosine using whole E. coli cells as a biocatalyst (Zaitseva et al. 1999) and a multitude of

other purine 2′-deoxy-2′-fluororibosides with antiviral activity using a combination of EcTP and EcPNP

as a biocatalyst (Tuttle et al. 1993).

However, dUrd2′F and dUrd2′F are very poor substrates in phosphorolysis reactions. This is presumably

a result of increased strength of the glycosyl bond as it follows from the crystallographic data for the

N1-C1′ bond length of uridine (average value 1.490 Å (Green et al. 1975) and its 2′-deoxyfluoro

analogues (1.454 Å (Marck et al. 1982) and 1.460 Å (Hempel et al. 1999), respectively). Moreover,

introduction of a fluorine atom into pentofuranose ring of nucleosides results in dramatic changes of

the conformation of such analogues precluding the formation of the productive substrate-catalytic

centre complex (for more detailed discussion, see (Mikhailopulo and Miroshnikov 2011)). Indeed, it

was reported that i) dUrd2′F showed no detectable substrate activity towards EcUP, ii) EcTP catalyzed

the phosphorolysis of dUrd2′F but at an extremely low rate and iii) the enzymatic cleavage of the

Enzyme Km (µM) kcat (s

-1) kcat/Km (s-1 µM-1)

Reference Uridine Thymidine Uridine Thymidine Uridine Thymidine

GtPyNP 2342 1282 275 83 0.12 0.06 This study

TtPyNP 145 435 279 679 1.92 1.56 This study

GsPyNP 190 460 - - - - (Hori et al. 1990)

EcTP 60 300 < 1·10-4 198 < 1.7·10-6 0.66 (Panova et al. 2004, Panova et al. 2007)

EcUP 80 270 98 5 1.22 0.02 (Alexeev et al. 2010, Panova et al. 2004)

Characterization of thermostable NPs 81

glycosidic bond of dUrd2′F equally afforded a high amount of enzyme and prolonged reaction time (6

days) (Tuttle and Krenitsky 1992, Tuttle et al. 1993).

In this study we have investigated the phosphorolysis of these challenging substrates by GtPyNP and

TtPyNP (Figure 38). Our results indicate that TtPyNP might be a good alternative to the use of E. coli

enzymes, but the use of GtPyNP is apparently not suitable for the applications discussed above: no

activity towards dUrd2′F and only poor activity towards dUrd2′F (0.44 % substrate conversion) was

detected with GtPyNP as biocatalyst after 30 min reaction time.

By contrast, the TtPyNP catalyzed reaction under the same conditions resulted in the phosphorolytic

cleavage of 0.65 % of dUrd2′F and 7.0 % of dUrd2′F. Since the optimal reaction temperature of TtPyNP

is significantly higher than 60 °C (section 4.1.2), we repeated the same reaction also at 80 °C. Now,

the TtPyNP catalyzed reaction resulted in 1.4 % phosphorolyzed dUrd2′Fand 15.6 % of dUrd2′F.

However, under these conditions the formation of two new peaks was observed by HPLC analysis of

the reaction mixture that contained dUrd2′F. The retention times coincide with those of authentic

samples of O2,2′-anhydro-1-(-D-arabinofuranosyl)uracil (anhydro-Urd) (2.3 min) and 1-(-D-

arabinofuranosyl)uracil (ara-U) (3.8 min). Hence, the formation of anhydro-Urd resulting from HF

release from the dUrd2′F molecule and the subsequent hydrolysis of anhydro-Urd resulting in ara-U

appear to be a reasonable explanation (Scheme 1). Phosphorolysis of both 2′-fluorosubstituted

pyrimidine nucleosides catalyzed by TtPyNP was also monitored over prolonged reaction times

(Figure 39A). The results show that the conversion of dUrd2′F at 80 °C could be increased to 46 % after

17 h; side-product formation was not observed. The final amount of phosphorolyzed dUrd2′F after

17 h at 80 °C (65 %) was in similar range as the amount obtained at 60 °C (60 %). By contrast, side

product formation, as discussed above, at 80 °C was significantly higher than at 60 °C (Figure 39B):

After 17 h 8.3 % of dUrd2′F reacted to anhydro-Urd at 80 °C, while the same value is decreased to

1.2 % at 60 °C.

O

HO

HO

F

HN

N

O

O

O

HO

HO

N

N

O

OO

NH

N

O

O

HO

HO

OH

-HF

Hydrolysis

O2,2'-Anhydro-(-D-ara-

binofuranosyl)uracil

1-(-D-arabinofuranosyl)-

uracil (ara-U)

dUrd2'F

Scheme 1

82 Characterization of thermostable NPs

0

10

20

30

40

50

60

70

80

90

100

Uridine Thymidine 2’Fana-U Urd2'F

Sub

stra

te c

on

vers

ion

[%

]

GtPyNP 60 C

TtPyNP 60 C

TtPyNP 80 C

GtPyNP 60 °CTtPyNP 60 °CTtPyNP 80 °C

O

HO OH

HO

NH

N

O

OO

HO

HO

NH

N

O

OO

HO

HO

NH

N

O

O

F

O

HO F

HO

NH

N

O

O

dUrd2'F dUrd2'F

Figure 38: The percentage of natural and 2′-fluorosubstituted pyrimidine nucleosides that were phosphorolytically cleaved by TtPyNP or GtPyNP after 30 min is shown. The reactions were performed with an enzyme loading of 0.1 mg ml-1 at the indicated temperatures. Figure adapted from (Szeker et al. 2012).

0

10

20

30

40

50

60

70

80

0 2 4 6 8 10 12 14 16 18 20

Sub

stra

te c

on

vers

ion

[%

]

Reaction time [h]

FU 80 °C

FU 60 °C

FanaU 80 °C

FanaU 60 °C

dUrd2‘F 80 °C

dUrd2‘F 60 °C

dUrd2‘F 80 °C

dUrd2‘F 60 °C

0

1

2

3

4

5

6

7

8

9

1.5 3 17

Sid

e p

rod

uct

co

nve

rsio

n [

%]

Reaction time [h]

Anhydro-Urd 60 °C

Anhydro-Urd 80 °C

Ara-U 80 °C

A) B)

Anhydro-Urd 60 °C

Anhydro-Urd 80 °C

Ara-U 80 °C

Figure 39: (A) Progress curves of the phosphorolysis reactions catalyzed by TtPyNP at 60 °C and 80 °C with dUrd2′F and dUrd2′F as substrates. (B) Side-product formation observed in the reactions with dUrd2′F as substrate. The percentage of dUrd2′F molecules that reacted to anhydro-Urd and ara-U during the TtPyNP catalyzed phosphorolytic cleavage reaction is shown. The ara-U formation at 60 °C could not be accurately determined but was estimated not to be higher than 0.4 %. Figure adapted from (Szeker et al. 2012).

Characterization of thermostable NPs 83

4.1.5. PNP activity of PyNPs

The previous section has shown that both PyNPs efficiently catalyze the phosphorolysis of the natural

pyrimidine nucleoside substrates uridine and thymidine. Considerably weaker activities were also

found for the 2′-fluorosubstituted uridine analogues.

In the further course of this study it was also recognized that not only pyrimidine nucleoside

analogues, but, unexpectedly also purine nucleosides are phosphorolyzed by both PyNPs. Initially we

considered it equivocal, whether this PNP activity indeed originates from the PyNPs themselves or

whether it could be the result of E. coli PNP contamination of the purified enzyme preparations. In

order to investigate this issue, the temperature optima of the apparent PNP activities of both PyNPs

were determined with inosine as substrates (Figure 40). The dependence of the reaction rate over

time is in accordance with the temperature dependence observed with the natural substrate uridine

(Figure 36A). These results clearly indicate that the PNP activity is inherent to the PyNPs and not the

result of a possible contamination with the PNP of the host (EcPNP).

0

20

40

60

80

100

20 30 40 50 60 70 80 90 100

Rel

ativ

e ac

tivi

ty [%

]

Temperature [°C]

GtPyNP

TtPyNP

Figure 40: Temperature dependence of the PNP activity of GtPyNP and TtPyNP. The highest reaction rate observed for each enzyme was set to 100 %. The purine nucleoside substrate was inosine.

To further characterize the PNP “side” activity of the PyNPs, estimated specific activities for different

purine and pyrimidine nucleoside substrates were determined (Table 13). The results show that

GtPyNP seems to have a preference for adenosine over inosine, while for TtPyNP no clear preference

can be seen. Even though the reaction rates for the purine nucleosides are extremely low in

comparison to the reaction rates observed with natural enzymes, they are significantly higher than

the reaction rates observed with the 2′-fluorosubstituted pyrimidine nucleosides. A simple conclusion

may be that the enzymes readily accept modifications on the heterocyclic base, but have stricter

requirements in respect to the sugar residue. These results prompted us to also investigate enzyme

activity towards cytidine, which only differs from uridine by one oxo-group replaced by an amino-

84 Characterization of thermostable NPs

group on the pyrimidine base. However, no phosphorolytic activity of GtPyNP and only poor activity

of TtPyNP were determined for cytidine as substrate. It appears quite remarkable that a single

substitution on the pyrimidine ring has such a pronounced impact on activity, while the substitution

of uracil by the purine bases adenine or hypoxanthine, results in still reasonable reaction rates.

Table 13. Estimated specific activities of GtPyNP and TtPyNP acting on pyrimidine and purine nucleosides.

Substrate Estimated specific activity [U mg-1 ]

GtPyNP (60 °C) TtPyNP (80 °C)

Uridine 51.51 148.80

Thymidine 22.44 286.00

Adenosine 4.57 × 10-1 8.09 × 10-1

Inosine 3.08 × 10-1 9.00 × 10-1

dUrd2′F 1.58× 10-3 2.62 × 10-2

dUrd2′F UDL 2.42 × 10-3

Cytidine UDL 8.7 × 10-4

Experiments performed with 1 mM substrate, at 60 °C (GtPyNP) or 80 °C (TtPyNP). UDL = under detection level

4.1.6. Characterization of thermostable PyNPs - summary and conclusions

PyNPs isolated from thermophilic microorganisms are promising biocatalysts for the efficient

synthesis of modified nucleosides (Taran et al. 2009). Up to now, the biocatalytic characterization of

PyNPs from thermophilic microbes was restricted to PyNPs from G. stearothermophilus strains

(Hamamoto et al. 1996, Hori et al. 1990, Saunders et al. 1969). We have studied here the

biocatalytical properties of two additional thermostable PyNP, originating from

G. thermoglucosidasius and T. thermophilus.

Our results indicate that both enzymes show excellent biocatalytical properties for applications with

natural pyrimidine nucleosides as substrates (thymidine, uridine) and a reaction temperature of

60 °C. In addition, the unusually high thermal stability of TtPyNP makes this biocatalyst also suitable

for reactions requiring an even higher reaction temperature of 80 °C.

We have further tested both thermostable PyNPs towards their ability to phosphorolyze

2′-fluorosubstituted pyrimidine nucleosides that have been shown to be very poor substrates in

phosphorolysis reactions employing EcUP or EcTP as biocatalyst (Tuttle and Krenitsky 1992, Tuttle et

al. 1993). Our results reveal striking differences of the substrate specificities of GtPyNP and TtPyNP,

in favour of the latter. The yield of phosphorolyzed 2′-fluorosubstituted pyrimidine nucleosides could

be enhanced by extending the reaction time and increasing the reaction temperature from 60 °C to

80 °C. However, the higher reaction temperature was not suitable for dUrd2′F since this substrate was

Characterization of thermostable NPs 85

not stable. By contrast, the higher reaction rate was a useful measure to increase the rate of the

phosphorolysis reaction of the more stable epimeric counterpart dUrd2′F.These findings make TtPyNP

a candidate as powerful biocatalyst in the transglycosylation reactions aiming at the synthesis of

2′-fluoro substituted purine nucleosides.

It was further discovered that both PyNPs studied here are able to accept also the purine nucleosides

inosine and adenosine as substrate. To our best knowledge PNP activity was so far only observed for

enzymes with the NP-I fold. PyNP however, belongs to the structurally unrelated NP-II family that is

likely to have evolved independently (Pugmire and Ealick 2002). The fact that purine nucleosides are

weakly accepted as substrates of PyNPs may rise the question whether PyNPs have diverged from

ancient enzymes with broader substrate specificity, also involved in the metabolism of purine

nucleosides.

The PNP “side activity” of the PyNPs could be of interest for synthetic reactions, because it could

allow to perform transglycosylation from a pyrimidine nucleoside donor to a purine base acceptor

with a single enzyme. In a similar way, NdRTs can be used for transglycosylations (see section 1.2.2).

However, these enzymes are specific for deoxyribosides and are therefore not suitable for the

synthesis of ribosides.

4.2. Thermostable enzymes with PNP activity

PNPs are thoroughly investigated enzymes. In fact, BRENDA (BRaunschweig ENzyme Database,

http://www.brenda-enzymes.org (Scheer et al. 2011)) currently lists PNPs from 73 different

organisms. Particularly the recognition of PNP as therapeutic target (for more details see section

1.3.1) has triggered a multitude of scientific studies devoted to the elucidation of structural

properties and catalytic mechanism (reviewed in (Bzowska et al. 2000, Silva et al. 2007)). Also, the

PNP-catalyzed reversible phosphorolysis of purine nucleosides has been recognized as an important

reaction that can be used for the synthesis of modified nucleosides (see section 1.2.3 for more

details). Therefore a number of PNPs have been investigated and subsequently exploited as

biocatalysts in transglycosylation reactions of nucleosides. Examples include the PNPs from E. coli

(Krenitsky et al. 1981), G. stearothermophilus strains (Hamamoto et al. 1996, Hamamoto et al. 1997a,

Hori et al. 1989b, Hori et al. 1991, Taran et al. 2009), Bacillus halodurans (Gordon et al. 2011, Visser

et al. 2010), and from A. hydrophila (Ubiali et al. 2012).

The use of NPs as biocatalysts in enzymatic transglycosylation reactions is also in focus of this work.

While in the last section the characterization of biocatalytical properties of thermostable PyNPs has

been presented, this section is devoted to thermostable enzymes with PNP activity. Specifically two

novel PNPs derived from the thermophilic microorganisms D. geothermalis and

G. thermoglucosidasius were investigated. In both microorganisms each two genes are annotated as

PNPs. As it will be discussed in more detail in the following section, high-molecular-mass-type PNPs

(hexameric form) were chosen to be studied here. In contrast to the low-molecular-mass (trimeric)

species, these PNPs are known not to be restricted to the phosphorolysis of 6-oxopurine nucleosides

and are therefore generally considered to have broader substrate specificity (see section 1.3.3).

86 Characterization of thermostable NPs

Moreover we investigated MTAP from the hyperthermophilic archaeon A. pernix. MTAPs from

hyperthermophilic sources have been successfully recombinantly expressed and structural features

have been thoroughly studied (Appleby et al. 2001, Cacciapuoti et al. 2005, Cacciapuoti et al. 2011).

Investigations of substrate specificities has shown that some MTAPs are able to accept both 6-oxo

and 6-aminopurine nucleosides as substrates (Cacciapuoti et al. 1994, Cacciapuoti et al. 2003), which

makes them interesting candidates as versatile catalysts for NP synthesis in our opinion. Indeed, a

recent study reports on the application of UP and PNP from A. pernix for the production of

5′-methyluridine (Zhu et al. 2012), whereby the coding region of the primers used for the isolation of

the thermostable PNP coincides with the coding sequence for ApMTAP. Hence the enzyme assigned

by the authors as PNP is most likely identical to ApMTAP here under investigation. However, in this

report (Zhu et al. 2012) no data concerning the biocatalytical properties of ApMTAP are disclosed. In

order to gain further knowledge concerning the potential application fields of ApMTAP as biocatalyst,

we therefore investigated the biocatalytical properties and substrate specificities toward natural and

artificial nucleosides.

4.2.1. Sequence analysis and homology modelling

The PNPs recombinantly expressed and characterized here are: PNP from D. geothermalis (GenBank

accession numbers ABF45792, YP_604961.1), PNP from G. thermoglucosidasius 11955 (sequence in

accordance with GenBank accession numbersEFG53380, AEH47728.1), and MTAP from A. pernix

(GenBank accession number NP_147653).

Phylogenetic analysis

A phylogenetic analysis with PNPs and MTAPs that are already reported in literature (Figure 41)

reveals that the PNPs from D. geothermalis and G. thermoglucosidasius studied here, show close

evolutionary relationships to other PNPs displaying the characteristics of hexameric, high molecular

mass PNPs (see section 1.3.3). These bacterial type PNPs are generally known to accept both,

6-oxopurine nucleosides (inosine, guanosine) and 6-aminopurine nucleosides (e.g. adenosine) as

substrate. Representatives are the 234 amino acid long PNP from G. stearothermophilus (Hamamoto

et al. 1997a), the 239 amino acid long PNP from E. coli (Jensen and Nygaard 1975) and the 238 amino

acid long PNP from A. hydrophila (Ubiali et al. 2012). Also the 235 amino acid long PNP from Bacillus

anthracis belongs to this group. This protein is identical to B. cereus adenosine phosphorylase that

was reported to prefer adenosine over 6-oxopurine nucleosides as substrate (Dessanti et al. 2012,

Sgarrella et al. 2007). The same applies to the 233 amino acid long PNP from B. subtilis (Jensen 1978),

although the preference for adenosine seems not so clear from the data of a more recent study (Xie

et al. 2011).

The phylogenetic analysis also illustrates that the second PNP of G. thermoglucosidasius (274 amino

acids) that was not examined in this study belongs to a group of PNPs found in Geobacilli and Bacilli

species, representing the eukaryotic (low molecular mass) type PNPs, implying specificity towards

6-oxopurine nucleosides, whereas 6-aminopurine nucleosides are not accepted as substrates.

Representatives of this group of PNPs are the 274 amino acid long PNP from G. stearothermophilus

Characterization of thermostable NPs 87

(Hamamoto et al. 1997b), the 272 amino acid long PNP from Bacillus halodurans (Visser et al. 2010),

and the 271 amino acid long PNP from B. subtilis(Jensen 1978, Xie et al. 2011). The second PNP from

D. geothermalis (238 amino acids) shows high sequence identity (61 % sequence identity) with the

235 amino acid long PNP from T. thermophilus which is in fact a hexameric protein but shows

substrate specificity as typical for trimeric (eukaryotic type) PNPs (Tahirov et al. 2004).

The MTAP from A. pernix cloned here (244 amino acids) appears to have an evolutionary relationship

to MTAPI from S. solfataricus (236 amino acids). The latter, has been shown to accept MTA and

adenosine, as well as 6-oxopurine nucleosides. By contrast, the MTAPII from S. solfataricus (270

amino acids), is specific for MTA and adenosine (Cacciapuoti et al. 2005) and appears closely related

to the second MTAP from A. pernix (275 amino acids) that is therefore catalytically less interesting

for this study and not investigated.

Figure 41: Phylogentic analysis of some PNPs and MTAPs described in literature. Enzymes are indicated by: Two-letter code for the source microorganism, enzyme name (PNP, MTAP), number of amino acids and the GenBank accession number. In the case of TtPNP the PDB code is given instead. Species abbreviations used: Gt = G. thermoglucosidasius, Gs = G. stearothermophilus, Bh = Bacillus halodurans, Ba = Bacillus anthracis, Bs = B. subtilis, Hs= Homo sapiens, Ec = E. coli, Ap = A. pernix, Ss = S. solfataricus, Pf = P. furiosus, Ah = A. hydrophila, Tt = T. thermophilus.

Homology modelling

The amino acid sequences of the PNPs investigated here (DgPNP, GtPNP) were blasted against the

amino acid sequences belonging to PNPs with resolved crystal structures (PDB database). Both

enzymes aligned best with the PNP from Bacillus anthracis (assigned here as BaPNP, PDB ID: 1XE3)

with 61 % (DgPNP) and 77 % (GtPNP) sequence identity, respectively. The PNP from Bacillus anthracis

(causative microorganism for anthrax) is identical to the B. cereus adenosine phosphorylase, for

88 Characterization of thermostable NPs

which the crystal structure was equally resolved (PDB ID: 2AC7, 3UAW) and investigated with respect

to the substrate specificity of the enzyme (Dessanti et al. 2012, Sgarrella et al. 2007). The crystal

structure of BaPNP was used as template to model DgPNP and GtPNP. The structural fold of the

models of DgPNP and GtPNP is almost identical to the template structure. As an example the

predicted structure of a monomer of GtPNP is shown, superimposed on the template BaPNP

hexamer (Figure 42A).

ApMTAP aligned best to the MTAP of S. solfataricus (PDB ID: 1JDS) with 45 % sequence identity and

an ApMTAP 3-dimensional model was built with SsMTAP as template. The S. solfataricus enzyme was

shown to be a hexamer with a broad substrate spectrum (Appleby et al. 2001, Cacciapuoti et al.

1994). Three intersubunit disulfide bonds link the dimers to each other to form a hexamer. In

SsMTAP the cysteine residues on position 125 are involved in this disulfide linkage. Since both in the

amino acid sequence alignment and the three dimensional superposition of SsMTAP and ApMTAP

Cys125 of SsMTAP corresponds to Cys129 of ApMTAP, it appears likely that in ApMTAP this cysteine

residue (on position 129) is equally involved in a disulfide linkage. This finding is in contradiction to

the predicted oxidation state of the cysteine residues summarized in Table 8, where Cys112 instead

of Cys129 was predicted to be in the bonding state. The crystal structure of SsMTAP together with

the superimposed monomeric model of ApMTAP is shown in (Figure 42B). However, a closer

investigation on the oligomerization state is necessary to determine the validity of this model, since

SDS-PAGE analysis of ApMTAP samples not treated with DTT may give evidence for a tetrameric

configuration (Figure 24).

GtPNP ApMTAP

Figure 42: GtPNP model based on BaPNP; ApMTAP model (with disulfide bridges) based on SsMTAP. The overall hexameric template structures as well as the superimposed monomeric models are shown. In the MTAP structure the disulfide bridges connecting each two dimers are highlighted (red).

4.2.2. Thermal characteristics

In order to assess the optimal temperature range at which the enzymes with PNP activity studied

here can be used in biocatalytic reactions, the optimal temperatures and thermal stabilities were

studied. The results are summarized in Table 14.

Characterization of thermostable NPs 89

Analysis of the reaction rates in dependence of the reaction temperature revealed that DgPNP shows

optimal activity at around 55 °C (Figure 43) which is slightly higher than the optimal growth

temperature of the source microorganism D. geothermalis (Ferreira et al. 1997). At 60 °C DgPNP

rapidly loses activity (stability half life approx. 1.6 h), while at 55 °C DgPNP is relatively stable over

prolonged incubation times (Figure 44). In comparison to the PNP from E. coli, DgPNP appears thus to

have a slightly lower temperature optimum, but is superior in regard to a higher stability at elevated

temperatures (Table 14).

Table 14. Thermal characteristics of enzymes with PNP activity from thermophilic microorganisms and E. coli.

Enzyme Organism Thermal stability (t1/2) Temp. optimum Reference

PNP* D. geothermalis > 8 h (55 °C)

1.7 h (60 °C) 55 °C This study

PNP* G. thermoglucosidasius > 8 h (70 °C)

6.3 h (75 °C) 70 °C This study

MTAP A. pernix > 27 h (90 °C) > 90 °C This study

PNP Bacillus halodurans 20.8 h (60 °C) 70 °C (Visser et al. 2010)

PNP I G. stearothermophilus - 70 °C (Hamamoto et al. 1996)

PNP II* G. stearothermophilus > 24 h ( 70 °C) 70 °C (Hamamoto et al. 1997a,

Taran et al. 2009)

PNP P. furiosus > 4 h (100 °C) 120 °C (Cacciapuoti et al. 2007)

MTAP P. furiosus > 5 h (100 °C)

43 min (130 °C) 125 °C (Cacciapuoti et al. 2003)

MTAP S. solfataricus > 2 h (100 °C)

15 min (130 °C) 120 °C (Cacciapuoti et al. 1994)

MTAPII S. solfataricus > 5 min (120 °C) 120 °C (Cacciapuoti et al. 2005)

PNP* E. coli > 30 min (50 °C)

< 30 min (55 °C) 60 °C (Li et al. 2008)

* bacterial, hexameric, high molecular mass type PNP

GtPNP showed optimal activity at 70 °C (Figure 43), which is about 10 °C higher than the temperature

optimum of GtPyNP (Table 11). Moreover, GtPNP is significantly more stable at higher temperature

in comparison to GtPyNP. At 70 °C no significant loss of activity can be seen within 8 h incubation

time (Figure 44), while stability half life of GtPyNP at 70 °C was only 1.6 h (Table 11). Similarly, the PNP

of E. coli was shown to be more thermoactive than E. coli thymidine phosphorylase, which can be

seen as the equivalent of PyNP in E. coli (Krenitsky et al. 1981). Likewise the PNPII was reported to

90 Characterization of thermostable NPs

be more thermostable than the PyNP in G. stearothermophilus (Taran et al. 2009). The PNP from

G. thermoglucosidasius here under investigation displays similar thermal properties as the PNPII from

G. stearothermophilus (Table 14). In fact, best to our knowledge, both enzymes represent the most

thermostable reported bacterial type PNPs that share a broader substrate specificity by accepting

both 6-oxopurine and 6-aminopurine nucleosides.

The reaction rates of ApMTAP increased with the reaction temperature up to the highest

temperature tested (99 °C, Figure 43). Furthermore, the enzyme appears to be extremely

thermostable with a stability half life exceeding 27 h at 90 °C (Figure 44). The high degree of

thermostability is conceivable, having in mind that the source microorganism is actually the

hyperthermophilic archaeon A. pernix, that shows optimal growth between 90 °C and 95 °C (Sako et

al. 1996). As the following section will show, ApMTAP together with the MTAPs of P. furiosus

(Cacciapuoti et al. 2007) and S. solfataricus (Cacciapuoti et al. 2005) thus belongs to the most

thermostable enzymes described to phosphorolyze adenosine as well as 6-oxopurine nucleosides

(Table 14). Both other enzymes have, however, not been investigated towards potential applications.

0

20

40

60

80

100

20 30 40 50 60 70 80 90 100

Rel

ativ

e ac

tivi

ty [%

]

Temperature [°C]

DgPNP

GtPNP

ApMTAP

Figure 43: Temperature optima of reaction rates. Relative activity of DgPNP, GtPNP,and ApMTAP over the reaction temperature assessed by the phosphorolysis of inosine. The highest reaction rate determined was set to 100 % for each enzyme.

As a conclusion for their application, it is not recommended to use DgPNP and GtPNP at

temperatures exceeding 55 °C and 70 °C, respectively, because both enzymes are rapidly inactivated

above these temperatures. By contrast, ApMTAP is highly thermostable and can be used even at

90 °C. The following chapter will show, that the reaction temperature of biocatalytic reactions

involving ApMTAP as catalyst will therefore not be restricted by the stability of the enzyme but

rather by the stability of substrates, intermediates, and products, or by the stability of the second

enzyme needed for the phosphorolysis of pyrimidine nucleosides.

Characterization of thermostable NPs 91

0

20

40

60

80

100

120

0 2 4 6 8

Res

idu

al a

ctiv

ity

[%]

Incubation time [h]

GtPNP 70 C

GtPNP 75 C

0

20

40

60

80

100

120

0 2 4 6 8

Res

idu

al a

ctiv

ity

[%]

Incubation time [h]

DgPNP 55 C

DgPNP 60 C

0

20

40

60

80

100

120

0 5 10 15 20 25 30

Res

idu

al a

ctiv

ity

[%]

Incubation time [h]

ApMTAP 90 C

°

°

°

°

°

Figure 44: Thermal stabilities of purine nucleoside phosphorolyzing enzymes. Protein samples were incubated for defined time intervals and subsequently the residual activity was determined. The activity of enzyme preparations that were not thermo-treated was set to 100 %.

4.2.3. Substrate specificities

The first aim was to determine which nucleosides are recognized as substrates by the thermostable

enzymes with PNP activity here under investigation. Therefore a number of natural and artificial

purine nucleoside analogues were screened towards their substrate activity. Enzyme concentrations

and sampling times were adjusted to ensure that the phosphorolysis reaction proceeded in linear

dependence of time and amount of NP added. The resulting apparent specific activities are

summarized in Table 15.

Phosphorolysis of natural purine nucleosides

All three enzymes recognize both inosine and adenosine as substrate. To compare the catalytical

properties with respect to recognition of adenine and hypoxanthine nucleosides, the relative

reaction rates were calculated from the data shown in Table 15 and listed together with the

corresponding rates of other enzymes reported in literature in Table 17. As pointed out before, PNPs

can be grouped in two categories: PNPs belonging to the low-molecular mass eukaryotic type

enzymes are specific for 6-oxopurine nucleosides (e.g. xanthosine phosphorylase from E. coli, PNPI

from G. stearothermophilus, PNP from B. subtilis and Mycobacterium smegmatis). PNPs categorized

as high-molecular mass, bacterial-type enzymes are generally characterized by a broader substrate

specificity and accept both 6-oxopurine nucleosides and 6-aminopurine nucleosides (e.g. the PNP of

E. coli encoded by the deoD gene, PNP encoded by the deoD gene of A. hydrophila, Ado-PNP from

Mycobacterium smegmatis, PNPII from G. stearothermophilus). Some members of the high-

molecular mass PNPs appear to have, however, a clear preference for adenosine over inosine as

substrate and are therefore also referred to as adenosine phosphorylases (examples listed in Table

17 are AdoP from B. cereus and from B. subtilis). The phylogenetic analysis illustrated in section 4.2.1

has shown that DgPNP and GtPNP are closely related to both the “normal” high-molecular mass PNPs

(e.g. PNPII from G. stearothermophilus, encoded by punB, E. coli PNP encoded by deoD, A. hydrophila

PNP encoded by deoD) as well as to the adenosine-specific nucleoside phosphorylases from Bacillus

92 Characterization of thermostable NPs

anthracis (enzyme is identical with AdoP from B. cereus) and B. subtilis (Figure 41). However, the

data summarized in Table 17 clearly indicate that from the catalytical properties both target PNPs

studied here behave similar to normal high-molecular mass PNPs that recognize both inosine and

adenosine efficiently as substrates.

Among the enzymes considered for the phylogenetic study, ApMTAP showed the closest

evolutionary relationship to the MTAPI from S. solfataricus (Figure 41). The similarity between both

amino acid sequences is also reflected by their catalytic properties. Both enzymes recognize

adenosine and inosine as substrate with a preference for the latter (Table 17).

Comparing the absolute activities for inosine determined in the present study with Vmax determined

by others for EcPNP (deoD) at 55 °C (211 U mg-1 (Li et al. 2008)) shows that the thermostable

enzymes investigated here appear to be less active. However, the results presented at this point are

possibly gained under nonsaturating substrate concentrations (1 mM) and therefore not necessarily

represent Vmax values. Moreover, it should be noted, that the PNPs investigated here are superior

with respect to thermostability. The data in Table 14 show that the stability half life of E. coli PNP at

55 °C is actually less than 30 min, while all enzymes recombinantly expressed in the present study are

stable at this temperature over prolonged incubation times. In comparison to AhPNP specific

activities (determined with 5 mM substrate concentration) (Ubiali et al. 2012), the activities

determined here are partially significantly higher, even though the substrate concentration was only

1 mM (Table 15). For PNPII of G. stearothermophilus a specific activity of 683 U mg-1 was reported

with adenosine as substrate at 70 °C (Hamamoto et al. 1997a). This value clearly exceeds the specific

activity determined here for GtPNP and the other enzymes. However, as stated above, the Vmaxvalues

(determined with saturating substrate concentrations) are possibly higher than the enzyme activities

obtained with 1 mM nucleoside concentration and listed for now in Table 15.

Table 15. Apparent specific activities obtained with a nucleoside substrate concentration of 1 mM.

Substrate Apparent specific activity [U mg-1]

DgPNP (55 °C) GtPNP (70 °C) ApMTAP (80 °C) AhPNP1 GsPNPII (70 °C)

2

Adenosine 151.88 379.16 41.27 35 683

Inosine 43.69 192.49 45.42 75 nd

dAdo2′F 0.023 0.009 0.013 nd nd

dAdo2′F 0.029 0.021 0.015 nd nd

Cytidine 0.0111 0.0147 0.0368 nd nd

1 = data from (Ubiali et al. 2012), were determined with 5 mM substrate concentration. 2 = data from (Hamamoto et al. 1997a). nd = not determined/ data not available; (*) 5 mM substrate concentration

In summary, the data presented here appear to be useful to categorize the target enzymes according

to their substrate specificity. Obviously all the enzymes show substrate specificities of the high

molecular mass (bacterial type) PNPs. The broader substrate specificity of these group of PNPs in

Characterization of thermostable NPs 93

comparison to the low molecular mass PNP is an important factor for possible future applications.

For example a halogen substitution on the C-2 position of the purine ring is tolerated only by high

molecular mass PNPs (Bzowska et al. 2000). On the other hand, kinetic studies will be needed to

investigate whether the differences in reaction rates observed here are due to substrate binding or

catalytic efficiency of the biocatalysts. Moreover, the maximal velocities and turnover number that

can be extracted from such a study will be more suitable determinants to compare the efficiency of

the thermostable PNPs among each other and with other reported enzymes.

Phosphorolysis of 2′-fluorosubstituted purine nucleoside analogues

Of particular interest for the present study is the synthesis of 2′-fluorosubstituted purine ribosides

from the respective pyrimidine nucleosides serving as pentofuranosyl donor and a purine base

serving as pentofuranosyl acceptor. In this transglycosylation reaction two types of NPs are

employed. The first enzyme, a PyNP phosphorolytically cleaves the donor nucleosides, while the

second enzyme with PNP activity is intended to catalyze the glycosyl bond formation between

pentofuranosyl-1-phosphate and the purine base (illustrated previously in Figure 6).The focus of this

section is to find out, which of the herein reported enzymes with PNP activity might be particularly

suitable for this application. For this purpose we have investigated the phosphorolysis of the two 2′-

fluorosubstituted purine nucleosides, 2′-deoxy-2′-fluoroadenosine (dAdo2′F) and 9-(2-deoxy-2-fluoro-

-D-arabinofuranosyl)adenine (dAdo2′F). The idea is that the enzyme that catalyzes most efficiently

the phosphorolysis of these substrates will also be the most suitable candidate for the reverse

reaction, required for the synthesis of the 2′-fluorosubstituted purine nucleosides.

The specific activities obtained with each 1 mM substrate for natural purine nucleosides and

2′-fluorosubstituted purine nucleosides are summarized in Table 15. The data show that the highest

absolute phosphorolysis rates for both compounds were obtained with DgPNP. By contrast, GtPNP

that actually appears to be the most efficient catalyst for natural purine nucleosides (inosine,

adenosine) shows less activity than DgPNP for dAdo2′F and is the weakest catalyst for dAdo2′F.

Table 16. Relative reaction rates (% of adenosine phosphorolysis).

dAdo2′F dAdo2′F

DgPNP (55 °C) 0.015 0.019

GtPNP (70 °C) 0.002 0.001

ApMTAP (80 °C) 0.032 0.036

1 mM, 50 mM KP, pH 7.0

This narrow specificity of GtPNP is also reflected by the extremely low relative reaction rates towards

the fluoro-substituted purine nucleosides shown in Table 16. Here the data of Table 15 describing the

phosphorolysis of fluoro-substituted purine nucleosides are expressed as relative reaction rates

whereby the reaction rate observed for adenosine phosphorolysis is set to 100 %. From these

94 Characterization of thermostable NPs

relative reaction rates one could also speculate that ApMTAP shows the broadest specificity. With

this enzyme dAdo2′F and dAdo2′F are phosphorolyzed with 0.032 % and 0.036 %of the reaction rate

observed with adenosine. By comparison, the reaction rate reported for EcPNP with 2′-deoxy-2′-

fluoroguanosine as substrate (1 mM) was 0.0031 % of the reaction rate with deoxyguanosine as

substrate (Tuttle et al. 1993).

Table 17. Relative reaction rates of enzymes with PNP activity for purine nucleosides.

Enzyme Source MO Relative reaction rates [%] Reference

Ino dIno Ado dAdo

PNP D. geothermalis 29 nd 100 nd This study

PNP G. thermoglucosidasius 51 nd 100 nd This study

AdoP/PNP B. cereus 2 2 100 46.7 (Sgarrella et al. 2007)1

AdoP/PNP B. subtilis 1 < 1 24 100 (Jensen 1978)2

PNPII (punB) G. stearothermophilus 46 40 100 90 (Hamamoto et al. 1997a)3

Ado-PNP Mycobacterium smegmatis 43 nd 100 58 (Buckoreelall et al. 2011)4

PNP (DeoD) E. coli 46 100 61 61 (Jensen and Nygaard 1975)

PNP (DeoD) A. hydrophila 55 100 26 20 (Ubiali et al. 2012)5

XanoP (XapA) E. coli 100 82 < 0.02 < 0.02 (Koszalka et al. 1988)6

PNPI (punA) G. stearothermophilus 100 100 0 nd (Hamamoto et al. 1996)

PNP B. subtilis 83 63 < 1 < 1 (Jensen 1978)6

PNP Mycobacterium smegmatis 100 nd - - (Buckoreelall et al. 2011)4

MTAP A. pernix 100 nd 91 nd This study

MTAPI S. solfataricus 100 nd 33 nd (Cacciapuoti et al. 2005)7

MTAP P. furiosus 36 nd 92 nd (Cacciapuoti et al. 2007)7

MTAPII S. solfataricus - nd 100 nd (Cacciapuoti et al. 2005)7

The reaction rates determined in this study where obtained with 1 mM nucleoside substrate concentration, 50 mM potassium phosphate buffer, pH 7.0 at 55 °C (DgPNP), 70 °C (GtPNP), and 80 °C (ApMTAP). 1) 150 µM nucleoside, 25 °C, 8.5 mM Na2HPO4, pH 7.5. 2) 1 mM nucleoside, 20 mM potassium arsenate, pH 7.1, 37 °C. For the PNP 100 % activity was assigned for the phosphorolysis of guanosine. 3) 20 mM nucleoside, 100 mM KP buffer, pH8, 70 °C. 4) 100 µM nucleoside, Ado-PNP had 137 % activity for 2-fluoroadenosine. 5) calculated from specific activities determined with 5 mM nucleoside substrate, 50 mM KP buffer, pH 7.5. 6) Numbers represent relative Vmax. 7) Numbers represent relative kcat values that were determined at 80 °C (P. furiosus enzymes) and 70 °C (S. solfataricus). For the MTAP from P. furiosus 100 % activity was assigned for the phosphorolysis of MTA. XanoP = abbreviation used here for xanthosine phosphorylase, the second PNP in E. coli encoded by the xapA gene that shows substrate specificity as eukaryoatic PNPs (specificity for 6-oxopurine nucleosides). (-) No activity was detected. (nd) = not determined/ data not available.

Characterization of thermostable NPs 95

Surprisingly, all 3 investigated enzymes seem to slightly better accept dAdo2′F than dAdo2′F as

substrate. By contrast, for the PyNPs there was a very pronounced preference for Urd2′F over Urd2′F

(see section 4.1.4).

The progress of the phosphorolysis reaction of dAdo2′F and dAdo2′Fwas also followed over a two hour

period (Figure 45).The results corroborate that DgPNP is indeed the best catalyst for phosphorolysis

of dAdo2′F. The phosphorolysis of dAdo2′F on the other hand is similar high for DgPNP and GtPNP.

0

5

10

15

20

25

0 0.5 1 1.5 2

dA

do

Fp

ho

sph

oro

lysi

s [%

]

Reaction time [h]

0

5

10

15

20

25

30

0 0.5 1 1.5 2

dA

do

Fp

ho

sph

oro

lysi

s [%

]

Reaction time [h]

ApMTAP 80 °C

DgPNP 55 °C

GtPNP 65 °C

A) B)

Figure 45: Phosphorolysis of dAdo2′F (A) and dAdo2′F (B) over time.

4.2.4. Characterization of thermostable PNP enzymes - summary and conclusions

Thermostable nucleoside phosphorylases are promising biocatalysts for enzymatic

transglycosylations aiming at the synthesis of modified purine nucleosides (Taran et al. 2009, Zhu et

al. 2012). The second step of the reaction scheme requires an enzyme with PNP activity. The high-

molecular mass “bacterial” type PNPs have the advantage of accepting both 6-aminopurine and 6-

oxopurine nucleosides and are therefore of particular interest for diverse synthetic applications.

However, up to now a closer inspection of biocatalytical properties and possible synthetic application

of high-molecular mass PNPs from thermophilic microorganisms is restricted to PNPII from

G. stearothermophilus strains (Hamamoto et al. 1997a, Hori et al. 1989a, Taran et al. 2009). We have

studied here some biocatalytical properties of two additional thermostable bacterial type PNPs,

which gene sequences were derived from G. thermoglucosidasius and D. geothermalis, respectively.

In addition we have studied MTAP from A. pernix as potential biocatalyst with PNP activity.

Homology modelling revealed that all three enzymes most probably have a hexameric configuration.

The characterization of the thermal properties shows significant differences for the temperature

optima (55 °C for DgPNP, 70 °C for GtPNP). For ApMTAP the reaction rates increased up to the

highest temperature tested (99 °C). Equally striking differences were found for the thermostability of

these enzymes. As concluded from the sequence alignments and from the catalytical properties the

PNPs appear to be reminiscent of other bacterial type PNPs (accepting both adenosine and inosine as

substrate) that have been described in literature as catalysts for the synthesis of nucleosides (PNPs

96 Characterization of thermostable NPs

from A. hydrophila (Ubiali et al. 2012), from G. stearothermophilus (Taran et al. 2009), and from

E. coli (Tuttle et al. 1993)). The advantage of all three biocatalysts in comparison to AhPNP and EcPNP

lies in the higher thermostability. From this point of view also ApMTAP is superior to GsPNP. We have

further investigated substrate specificities towards 2′-fluorosubstituted nucleosides. The idea was

that the enzyme most suitable for the phosphorolysis of 2′-fluorosubstituted purine nucleosides will

also be most suitable one for the reverse (synthetic) reaction. For dAdo2′F phosphorolysis, DgPNP

proved to be the best catalyst due to two criteria. Firstly, the highest reaction rate was observed with

DgPNP. Secondly, in the final application with dUrd2′F as donor the temperature should not exceed

60°C, since the donor is thermosensitive (see section 4.1.4). The phosphorolysis of dAdo2′F, however,

is similar high for DgPNP and GtPNP. Taking into account that the donor dUrd2′F is quite stable and

significant higher phosphorolysis rates were achieved with TtPyNP at 80°C than at 60 °C, GtPNP

appears to be superior over DgPNP as second enzyme in the transglycosylation reaction. However,

for this application also ApMTAP is possibly a good alternative due to its high stability at 80 °C.

These findings render all three enzymes promising for biocatalytic applications aiming at the

synthesis of 2′-fluorosubtituted nucleosides. The next chapter will focus on transglycosylations

employing PyNPs and enzymes with PNP activity at the same time and show, whether the combined

use of the generated thermostable enzymes will translate in high yield of synthetic purine

nucleosides. Moreover, the combination of both enzymes in transglycosylation reactions will also

allow to further characterize the substrate specificities of the PNPs towards artificial heterocyclic

bases, for which corresponding nucleosides were not readily available to investigate the reverse,

phosphorolytic reaction here.

5. Enzymatic transglycosylations with thermostable NPs

In this chapter the use of the generated thermostable biocatalysts for the synthesis of modified

purine nucleosides will be investigated. Our interest concerns the synthesis of sugar modified purine

nucleosides (2′-fluorinated nucleosides) and the synthesis of purine nucleosides with modified

nucleobases (2,6-dihalogenated purine nucleoside).

5.1. Introduction

5.1.1. Chemical synthesis of 2′-fluorinated nucleosides

The small size and high electronegativity renders fluorine a remarkable element. The C-F bond is one

of the strongest known bonds and organofluorine compounds are hence often characterized by high

chemical stability. On the other hand the bioisosteric replacements with fluorine have lead to the

disclosure of drugs with enhanced bioavailability, metabolic stability, and biological activity

(Hagmann 2008, Müller et al. 2007). Fluorination is therefore considered as an important tool in drug

discovery (Rentmeister et al. 2009).

The 2′ position of the carbohydrate moiety is the distinguishing feature between ribo- and

deoxyribonucleosides and is an attractive target for the selective introduction of a fluorine atom.

Here, fluorine can function as isosteric replacement of the hydrogen atom or as an isopolar mimic of

the hydroxyl group, respectively, whereby the C-F bond length (1.41 Å) is much closer to the C-O

bond length (1.35 Å) than to the C-H bond length (1.09 Å) (Liu et al. 2008, Müller et al. 2007).

Fluorine substitutions at the 2′ position of nucleosides have been shown to contribute to chemical

stability, in particular in acidic environment and lead to increased metabolic stability ((Liu et al. 2008)

and references therein).

Generally two strategies towards the synthesis of 2′-fluorinated nucleosides have been disclosed: i)

the direct fluorination of nucleosides and ii) the convergent synthesis that involves the condensation

of a fluorine-substituted ribose with a heterocyclic base. While the first approach may represent a

very efficient synthetic route for specific cases, the second approach is more versatile.

2′-Deoxy-2′-fluoro ribonucleosides

The efficient synthesis of 2′-deoxy-2′-fluorouridine starting from the natural ribonucleoside uridine

was already reported by Codington et al. in 1964. The approach that became the standard method

for fluorination of pyrimidine nucleosides involves the synthesis of O2,2′-anhydro-1-(-D-

arabinofuranosyl)uracil as intermediate product followed by a nucleophilic fluorination through

treatment with hydrogen fluoride. To avoid hazardous hydrogen fluoride, Olah’s reagent (mixture of

hydrogen fluoride and pyridine) has later been successfully employed as alternative fluorinating

reagent (Liu et al. 2008, Shi et al. 2005).

An analogous reaction route for the synthesis of 2′-deoxy-2′-fluoro ribofuranosyl purines does not

exist due to a lack of a suitable oxo-group on the heterocyclic base that could participate with the C-

98 Enzymatic transglycosylations with thermostable NPs

2′ atom in an anhydro bridge formation. Alternatively the synthesis of 2′-deoxy-2′-fluoroadenosine

and 2′-deoxy-2′-fluoroguanosine has been described by a nucleophilic displacement of a

corresponding triflate at the 2′-position (Kawasaki et al. 1993, Ranganathan 1977). However, due to

the number of different steps, the overall yield is limited. Furthermore it should be noted, that the

synthesis of the starting material (e.g. ara-A) also entails a number synthetic steps or chemo-

enzymatic approaches (Roshevskaia et al. 1986).

Another disadvantage of the described method is that it is not easily adaptable to the synthesis of

other 2′-deoxy-2′-fluoro ribofuranosyl purines. Therefore Thomas et al. investigated the convergent

synthesis by coupling 3,5-di-O-benzoyl-2-deoxy-2-fluoro-D-ribofuranosyl bromide to

2,6-dichloropurine (Thomas et al. 1994). The product could be readily converted to other analogues

resulting in the 2-fluoroadenine, 2-chloroadenine, 2,6 diaminopurine and guanine congeners by

standard procedures. The drawback of this approach is the formation of stereoisomers (α- and β-

anomers) and the need to separate both forms. Conversely, a chemo-enzymatic approach as it is

followed in the present work exclusively results in the biological active β-anomers and reduces the

number of required catalytic steps.

2′-Deoxy-2′-fluoro arabinonucleosides

While protocols for the efficient synthesis of 2′-deoxy-2′-fluoro ribonucleosides via direct fluorination

of preformed pyrimidine nucleosides have been developed, the synthesis of the according 2′-deoxy-

2′-fluoro arabinofuranosyl pyrimidine nucleosides is significantly more challenging. In fact the

susceptibility of pyrimidine nucleosides to form O2,2′-anhydro-bonds that is actually exploited for the

synthesis of pyrimidine 2′-deoxy-2′-fluoro ribonucleosides, is somehow precluding the efficient

synthesis of the respective arabinofuranosyl analogues via a direct fluorination approach. The

problem is that prior to the fluorination at the 2′-arabino position, the C2 carbonyl group of the

pyrimidine base will displace the leaving group at the C2-ribo activated function and O2,2′-anhydro-1-

(-D-arabinofuranosyl)pyrimidines is formed (Watts and Damha 2008). A subsequent SN2 type

reaction that entails the inversion of the configuration will then lead to the 2′-substituted

ribofuranosyl pyrimidine.

Therefore, the convergent synthetic approach towards 2′-deoxy-2′-fluoro arabinofuranosyl

pyrimidines has been in focus of research, and efficient protocols have been established. In 1979

Watanabe and co-workers described the synthesis of a series of 2′-deoxy-2′-fluoro arabinofuranosyl

pyrimidines with antiviral activity (Watanabe et al. 1979). Key steps of the synthetic procedure are

the fluorination and subsequent bromination of a protected ribofuranose derivative, and the

coupling with silylated pyrimidine. The protocol was further improved by Tann and co-workers in

1985 by proposing an efficient method for the selective synthesis of the respective 1-α-bromo sugar

(Tann et al. 1985). The advantage is that the 1-α-bromo sugar, in contrast to the stereoisomeric 1-β-

bromo sugar leads to preferred synthesis of (natural) β-nucleosides. Hence, 2′-deoxy-2′-fluoro

arabinofuranosyl pyrimidine can be prepared stereoselectively in high overall yield (Howell et al.

1988).

Enzymatic transglycosylations with thermostable NPs 99

On the other hand, the synthesis of the purine nucleoside congeners is significantly more complicate.

The major obstacle is the formation of stereo- and regioisomers (e.g. 7- or 9-substituted α- and

β-anomers) within the condensation of the ribose and purine base moiety, that entails the need of

tedious purification steps and eventually decreases the overall yield (Montgomery et al. 1986,

Tennilä et al. 2000, Wright et al. 1969). Despite impressive achievements in the development of

synthetic procedures for specific compounds, as for example for clofarabine (Bauta et al. 2004), the

synthesis of 2′-deoxy-2′-fluoro arabinofuranosyl purines remains in general a complicate endeavour.

Owing to these difficulties, alternative synthetic routes towards 2′-deoxy-2′-fluoro arabinofuranosyl

purines have been investigated. Thus, dAdo2′F and 9-(2-deoxy-2-fluoro--D-

arabinofuranosyl)hypoxanthine have been synthesized by direct fluorination of preformed

nucleosides (Krzeminski et al. 1991, Maruyama et al. 1999, Sivets et al. 2006). The fluorination step

itself was reported as fairly efficient (approx. 30 %). However, the overall synthesis is an advanced

multistep procedure involving the need of selective protection steps and is not comparable to the

facile and efficient synthesis of 2′-deoxy-2′-fluoro ribofuranosyl pyrimidines via direct fluorination

(described before).

Noteworthy, despite the aforementioned difficulties arising from the C2 carbonyl group, the direct

fluorination at the 2′-arabino position of pyrimidine nucleosides has been recently proposed

(Turkman et al. 2010). The procedure was applied for the synthesis of radiolabelled 2′-deoxy-2′-

fluoro-5-methyl-1-β-D-arabinofuranosyluracil ([18F]FMAU).

5.1.2. The chemo-enzymatic synthesis of 2′-fluorinated purine nucleosides

Motivated by the difficulties arising within the chemical synthesis of purine nucleoside analogues,

enzyme-assisted synthetic routes towards 2′-fluorinated purine nucleosides have been investigated.

Hereby, the enzymatic transfer of a 2′-fluorinated pentofuranose moiety from a pyrimidine

nucleoside donor to a purine base acceptor represents an attractive approach. With this strategy

Tuttle and co-workers synthesized 2′-deoxy-2′-fluoro ribofuranosyl purines employing 2′-deoxy-2′-

fluorouridine as pentofuranosyl donor (Tuttle et al. 1993). The enzymatic transglycosylation was

mediated by large amounts of EcTP and EcPNP and required long reaction times (up to 57 days). In

order to improve the availability of poorly soluble purine bases and to increase the efficiency of the

reaction, the temperature was increased from 37 °C to 50 °C. To retain the activity of the E. coli NPs,

enzymes were immobilized. Moreover, the first 24 h were run at 37 °C, in order to accumulate

pentose-1-phosphate that is known to further stabilize the NPs. The results demonstrate that this

“high” temperature approach was a clear advantage with respect to reduced reaction times and

increased yields.

The same strategy was patented for the synthesis of a number of 9-(2-deoxy-2-fluoro-β-D-

arabinofuranosyl)purines (Tuttle and Krenitsky 1992), whereby 1-(2-deoxy-2-fluoro-β-D-

arabinofuranosyl)thymine served as pentofuranosyl donor. Although the high-temperature approach

(at 50 °C with immobilized E. coli enzymes) was used, enzyme loadings were high and the reaction

time fairly long (e.g. 6 days for 2,6-Diamino-9-(2-deoxy-2-fluoro-β-D-arabinofuranosyl)-9H-purine).

100 Enzymatic transglycosylations with thermostable NPs

The enzymatic synthesis of 2′-deoxy-2′-fluoroguanosine with 2′-deoxy-2′-fluorouridine or -cytidine as

pentofuranosyl donors was also accomplished by the use of whole cells of E. coli BMT-4D/lA as

biocatalyst (Zaitseva et al. 1999).

Recently the possibility to use NdRTs for the transfer of 2′-deoxy-2′-fluoro ribofuranosyl moieties was

disclosed (Fernandez-Lucas et al. 2010). With dUrd2′F as donor and adenine as acceptor, dAdo2′F was

synthesized employing Lactobacillus reuteri NdRT

A modified approach was followed by Yamada and co-workers (Yamada et al. 2009). Here the

pentofuranosyl moiety was not provided by the phosphorolysis of a corresponding pyrimidine

nucleoside. Instead, the authors successfully synthesized 2-deoxy-2-fluoro arabinofuranosyl-α-1-

phosphate that was afterwards enzymatically linked to a purine base affording the desired 9-(2-

deoxy-2-fluoro-β-D-arabinofuranosyl)purine. In order to perform the reaction at 50 °C, thermostable

PNP from G. stearothermophilus was used. The approach is motivated by the fact that the

phosphorolysis reaction of the pyrimidine donor, that proceeds very slowly if the 2′ position is

substituted with a fluorine atom in the arabino position, can be avoided.

5.1.3. 2,6-Dihalogenated purine nucleosides

Purine nucleosides with variable substitutions on position 2 and 6 of the heterocyclic base are of

pharmaceutical interest due to antimicrobial and anticancer activities (Bellezza et al. 2008, Bonate et

al. 2006, Cappellacci et al. 2011, Rodenko et al. 2007). Insights into the biological activities of

adenosine receptors have further expanded potential therapeutic application fields (Nair et al. 1995,

Poulsen and Quinn 1998, Samsel and Dzierzbicka 2011, Vittori et al. 2000).

2,6-Dichloropurine nucleosides are valuable precursors for this class of nucleoside derivatives (e.g.

(Kazimierczuk et al. 1984, Montgomery et al. 1986, Tennilä et al. 2000). The chlorine atoms display

electron withdrawing centres on the purine ring, making these positions amenable for nucleophilic

substitutions. Hereby the chlorine atom on ring position 6 is much more reactive than the chlorine

atom at position 2 (Dobak et al. 2008, Schaeffer and Thomas 1958). This difference in reactivity

enables the synthesis of 2,6-substituted derivatives either by selective monosubstitution of the C-6

chlorine atom or by consecutive substitutions of the C-6 and C-2 chlorine atoms by diverse functional

groups (Salvatori et al. 2002, Schaeffer and Thomas 1958, Tennilä et al. 2000, Vittori et al. 2000,

Wright et al. 1987). Likewise 2,6-dichloropurine nucleosides can serve as starting material for cross-

coupling reactions leading to the efficient synthesis of 2-substituted 6-methylpurine nucleosides

(Hocek and Dvorakova 2003).

Simple and efficient methods for the preparation of 2,6-dichloropurine have been developed, for

example the facile synthesis from xanthine has been described (Zeng et al. 2004). By contrast, the

coupling reaction to carbohydrate moieties gave rise to regioisomers and, moreover, mixtures of the

α- and β-anomers in the case of 2′-deoxynucleosides (Vorbrüggen and Ruh-Pohlenz 2001). By strictly

obeying to regio- and stereoselective requirements, enzymatically catalyzed coupling reaction of 2,6-

dichloropurine to pentofuranosyl moieties could therefore present an attractive alternative.

Enzymatic transglycosylations with thermostable NPs 101

5.2. Synthesis of 2′-fluorosubstituted purine nucleosides

The aim was to synthesize 2′-deox-2′-fluoroadenosine (dAdo2′F) and 9-(2-deoxy-2-fluoro--D-

arabinofuranosyl)adenine (dAdo2′F) by enzymatic transglycosylations employing 2′-deoxy-2′-

fluorouridine (dUrd2′F) or 1-(2-deoxy-2-fluoro--D-arabinofuranosyl)uracil (dUrd2′F) as pentofuranosyl

donor and adenine as pentofuranosyl acceptor.

+ Pi+ Pi

Uracil

PyNPPNP/MTAP

NH

NH

O

O

O

HO F

HO

OPO32-

O

HO F

HO

N

NN

N

NH2

O

HO F

HO

NH

N

O

O

O

HO

F

HO

NH

N

O

O O

HO

F

HO

N

NN

N

NH2

O

HO

F

HO

OPO32-

dUrd2‘F

dUrd2‘F

Adenine

N

NNH

N

NH2

dAdo2‘F

dAdo2‘F

Figure 46: Enzymatic synthesis of dAdo2′F and dAdo2′F from respectively sugar-modified uridine analogues (dUrd2′F, dUrd2′F) and adenine, mediated by PyNP and PNP or MTAP.

5.2.1. Synthesis of 2′-deoxy-2′-fluoroadenosine

The experiments presented in section 4.1.4 indicated that TtPyNP compared to GtPyNP is clearly the

more efficient catalysts for the phosphorolysis of dUrd2′F. Even at 60 °C - where GtPyNP shows

optimal activity, but TtPyNP only about 12 % of the activity displayed at 80 °C (Figure 36)- the TtPyNP

catalyzed phosphorolysis of dUrd2′F is about 16 times more efficient than the GtPyNP catalyzed

reaction. On the other hand, in section 4.2.3 it was shown that among the tested enzymes with PNP

activity DgPNP is most efficiently catalyzing the phosphorolysis of dAdo2′F (determined at 55 °C),

followed by ApMTAP (at 80 °C). However, due to the themolability of the donor nucleoside (dUrd2′F)

discussed in section 4.1.4, the final reaction should not be conducted at a temperature considerably

exceeding 60 °C. This restriction makes the use of ApMTAP rather unattractive since at 60 °C this

highly thermoactive biocatalyst shows only about 22 % of the activity displayed at 80 °C (Figure 43).

For these reasons we tested the following two combinations of enzymes for the synthesis of dAdo2′F:

i) TtPyNP + DgPNP at 55 °C and ii) TtPyNP + GtPNP at 65 °C. Reactions were stopped after 2 and 18 h,

respectively. The formation of dAdo2′F was confirmed by comparing the retention time and UV

spectrum of the newly evolved peak with that of authentic sample of dAdo2′F. From the peak areas

obtained by HPLC measurements, the fraction of phosphorolytically cleaved donor (dUrd2′F)

molecules and the fraction of the acceptor (adenine) molecules that were transformed into the

102 Enzymatic transglycosylations with thermostable NPs

product (dAdo2′F ) were calculated (Figure 47). The second value can also be understood as the yield

with respect to the applied acceptor molecule.

0

10

20

30

40

18 h 2 h 18 h 2 h 18 h

TtPyNP 65 °C TtPyNP+GtPNP 65 °C TtPyNP+DgPNP 55 °C

Sub

stra

te c

on

vers

ion

%

phosphorylized FU

FA formation

Uracil/(dUrd2‘F + Uracil) × 100 %

dAdo2‘F/(Adenine + dAdo2‘F) × 100 %

Figure 47: Synthesis of 2′-deoxy-2′-fluoroadenosine (dAdo2′F) from 2′-deoxy-2′-fluorouridine serving as donor and adenine serving as pentofuranosyl acceptor.

The results show that the yield was in fact in similar range for both enzyme combinations tested and

that prolonged incubation times are needed to obtain a reasonable outcome (after 18 h the yield was

less than 20 %). When only TtPyNP was added to the reaction mixture the phosphorolysis reaction of

the donor molecule proceeded as expected while product formation was negligible (Figure 47).

When only PNPs were added, no substrate conversion was observed. In the reaction employing

TtPyNP and GtPNP at 65 °C, after 18 h 3.4 % of the donor molecules reacted to the unwanted side-

product O2,2′-anhydro-1-(-D-arabinofuranosyl)uracil (anhydro-Urd). In the TtPyNP-DgPNP catalyzed

reaction at 55 °C anhydro-Urd formation was significantly lower (2 %). Therefore the TtPyNP-DgPNP

catalyzed reaction was considered as more suitable and was further investigated with respect to the

reaction time (Figure 48). The formation of the product (dAdo2′F) increased in linear relation to the

reaction time, while the phosphorolysis of dUrd2′F appeared to follow a hyperbolic function over the

reaction time. Therefore it seems likely that the second reaction represents the rate-limiting step.

The final yield obtained after 24 h was 24 %. At this final point 44 % of the donor molecules were

phosphorolytically cleaved to uracil and the pentose-1-phosphate intermediate. Taking into account

that the initial donor nucleoside concentration was 2 mM and the initial acceptor concentration was

1 mM, these data show that in the 1 ml reaction vessel used, about 0.88 µmol pentose-1-phosphate

intermediate was formed, and 0.24 µmol of the intermediate further reacted to the final product.

Overall 2.3 % of the donor molecules reacted to the unwanted side-product anhydro-Urd. 1-(-D-

Arabinofuranosyl)uracil (ara-U) was not detected.

In summary, the combined use of TtPyNP and DgPNP at 55 °C was found to be a good approach for

the synthesis of dAdo2′F. The data suggest that the current product yield (23 % with respect to the

pentofuranosyl acceptor) is restricted by the reaction rates of the biocatalysts. Hence, the yield could

Enzymatic transglycosylations with thermostable NPs 103

be easily increased simply be extending the reaction time or increasing the enzyme loading. Other

factors that could be considered for future optimization include the molar concentration of

substrates and buffer, as well as the ratio of the TtPyNP and DgPNP enzyme loading.

44

23

0

10

20

30

40

50

0 5 10 15 20 25

Sub

stra

te c

on

vers

ion

[%

]

Reaction time [h]

phosphorylized FU

FA formation

Uracil/(dUrd2‘F + Uracil) × 100 %

dAdo2‘F/(Adenine + dAdo2‘F) × 100 %

Figure 48: Progress curves of the enzymatic transglycosylation reaction employing dUrd2′F as pentofuranosyl donor, adenine as acceptor, resulting in the synthesis of dAdo2′F. The reaction was mediated by TtPyNP and DgPNP at 55 °C.

Comparison to literature results

Synthesis of dAdo2′F by enzymatic transglycosylation with dUrd2′F as donor and adenine as acceptor

was described by Tuttle and co-workers who employed EcPNP and EcTP as biocatalysts (Tuttle et al.

1993) and, more recently, by Fernández-Lucas et al. who made use of the NdRT from

Lactobacillus reuteri (Fernandez-Lucas et al. 2010). As the following discussion will show, a direct

comparison of the efficiencies of the reactions is not always straight-forward, making it difficult to

judge immediately whether the here generated thermostable NPs represent superior biocatalysts.

For example the ratio of the substrates that were used in the reactions is one factor that precludes a

direct comparison of the results. In the E. coli NP-catalyzed transglycosylation reported by Tuttel et

al. the concentration of the acceptor molecule (adenine) was 2.5 times higher than the concentration

of the donor molecule (dUrd2′F), and the yield (82 % after 17 days) was calculated with respect to the

amount of dUrd2′F used in the reaction. By contrast in the present study 2 times more donor than

acceptor was employed and a yield of 23 % (with respect to pentofuranosyl acceptor) was reached

after only 1 day.

Another factor that poses difficulties for the comparison of the efficiency of the enzymes represents

the enzyme loading. Tuttle et al. express the enzyme loading in U, defined as the enzyme amount

that converts 1 µmol natural substrate (e.g. inosine) per min at 25 °C. Defining the amount of

thermostable enzymes in the same way would be misleading, since the activity at 25 °C would be

extremely low. Therefore the amount of enzyme added was not expressed in U but was instead

104 Enzymatic transglycosylations with thermostable NPs

simply set to 0.1 mg ml-1 in the present study. However, if the specific activities of the NPs are

known, it is obviously possible to calculate how much enzyme (in mg ml-1) was used in the study of

Tuttle and co-workers. For example it was stated that 3900 U EcPNP was employed in a reaction

volume of 20 ml (at least in the first 6 days). A maximal velocity of 90.5 U mg-1 (Li et al. 2008) and

230 U mg-1 (Jensen and Nygaard 1975) was reported for EcPNP at 37 °C. Since the temperature

optimum of EcPNP is 60 °C (Li et al. 2008) it can hence be speculated that the specific activity at 25 °C

is significantly lower than 230 U mg-1. Therefore it can be concluded that the enzyme loading of

EcPNP was considerably higher than 0.85 mg ml-1 within the first 6 days, whereas in the present

study only 0.1 mg ml-1 DgPNP was used.

As stated above, a direct comparison of the efficiencies of the enzymes is not possible with the

current data. However, taking the information on yield, reaction time and enzyme loading together,

there seems to be strong evidence that the efficiency of DgPNP and TtPyNP could considerably

exceed the efficiency of the E. coli NPs employed previously (Tuttle et al. 1993) for the synthesis of

dAdo2′F.

Furthermore, the here generated thermostable enzymes appear to be also superior to NdRT from

Lactobacillus reuteri, that was recently described as biocatalyst for the synthesis of dAdo2′F. The

reaction was performed at 40 °C with each 1 mM substrate concentration and an enzyme loading of

0.34 µg per 40 µl, which corresponds to 0.0085 mg ml-1. The specific activity assessed for dAdo2′F

synthesis after 24 h was 0.3 x 10-3 U mg-1. Note that there is a typo in the exponent in the publication

(Fernandez-Lucas et al. 2010). According to our calculations, if these data are transformed into yield

with respect to the amount of adenine this would correspond to approximately 3.8 x 10-4 %. By

contrast, in the present study DgPNP and TtPyNP (enzyme loading each 0.1 mg ml-1) were employed

at 55 °C with 2 mM dUrd2′F and 1 mM adenine, resulting in a yield of 23 % with respect to adenine.

These data indicate that, even if both experiments would have been performed with the same

enzyme loading, the final yield of dAdo2′F obtained with DgPNP and TtPyNP as biocatalyst would

considerably exceed the yield that would have been obtained with Lactobacillus reuteri NdRT.

5.2.2. Synthesis of 9-(2-deoxy-2-fluoro--D-arabinofuranosyl)adenine

After the successful synthesis of dAdo2′F the aim was now to synthesize the epimeric counterpart 9-

(2-deoxy-2-fluoro--D-arabinofuranosyl)adenine (dAdo2′F). Therefore dUrd2′F should serve as

pentofuranosyl donor and, again, adenine should serve as pentofuranosyl acceptor.

TtPyNP was again used as PyNP, because only this enzyme recognized dUrd2′F as a substrate, while

for GtPyNP no substrate activity could be seen (for further details see 4.1.4). Nevertheless, it should

be noted that also for TtPyNP the reaction rate observed with dUrd2′F as substrate was about 10

times lower than that recorded for dUrd2′F.

According to their phosphorolytic activity on dAdo2′F, DgPNP (at 55 °C) and GtPNP (at 70 °C) appear to

be the most efficient catalysts that could be used for dAdo2′F synthesis. Despite these findings we

decided to test the following two combinations first: i) TtPyNP + GtPNP at 70 °C (temperature is

limited by GtPNP stability) and ii) TtPyNP + ApMTAP at 80 °C (temperature is limited by TtPyNP

Enzymatic transglycosylations with thermostable NPs 105

stability). The use of DgPNP was not tested because this would limit the reaction temperature to

55 °C where the first, TtPyNP-catalyzed phosphorolysis of dUrd2′F would presumably proceed

unreasonably slow. While for dAdo2′F synthesis the reaction temperature was also limited by the

instability of the donor nucleoside dUrd2′F at temperatures above 60 °C, dUrd2′Fappeared to be stable

at even 80 °C and is therefore not limiting the reaction temperature.

0.37 0.48 0.650

5

10

15

20

25

30

18 h 2 h 2 h

TtPyNP 65 °C TtPyNP + GtPNP 70 °C TtPyNP + ApMTAP 80 °C

Sub

stra

te c

on

vers

ion

%

Series1

Series2

Uracil/(dUrd2‘F + Uracil) × 100 %

dAdo2‘F/(Adenine + dAdo2‘F) × 100 %

Figure 49: Biocatalyst screening for the synthesis of dAdo2′F by an enzymatic transglycosylation reaction employing dUrd2′F as donor and adenine as pentofuranosyl acceptor.

The procedure of the experiments was essentially the same as described for the synthesis of dAdo2′F

in the previous section. In a first step the enzymatic transglycosylation was investigated with both

the aforementioned enzyme combinations after a reaction time of 2 h (Figure 49). The yield of

dAdo2′F with respect to adenine was in a similar range (0.48 % and 0.65 %). If TtPyNP was employed

as sole biocatalyst the yield was only 0.37 % after 18 h. Hence TtPyNP alone was not further

investigated as biocatalyst for dAdo2′F synthesis. If PNPs were employed as sole biocatalysts no

substrate conversion could be seen (data not shown). In a next step the enzymatic transglycosylation

reaction affording dAdo2′F was monitored over time (Figure 50). The results indicate that the

combination of TtPyNP and ApMTAP at 80 °C was clearly superior over the combination of TtPyNP

and GtPNP at 70 °C, resulting in 24 % yield of dAdo2′F with respect to adenine after a reaction period

of 24 h.

Comparison to literature results

Tuttle and Krenitsky have patented the enzymatic synthesis of a number of 9-(2-deoxy-2-fluoro-β-D-

arabinofuranosyl)purines, whereby similarly as described for the synthesis of purine 2′-deoxy-2′-

fluororibosides the procedure involved the use of EcPNP and EcTP as biocatalysts (Tuttle and

Krenitsky 1992). The enzymatic transglycosylation reaction described in the present study followed

the same principle, with the exception that dUrd2′F instead of 1-(2-deoxy-2-fluoro-β-D-

arabinofuranosyl)thymine was used as pentofuranosyl donor.

106 Enzymatic transglycosylations with thermostable NPs

37

24

0

10

20

30

40

50

0 5 10 15 20 25

Sub

stra

te c

on

vers

ion

[%

]Reaction time [h]

Phosphorylized FanaU

FanaA formation

32

14

0

10

20

30

40

50

0 5 10 15 20 25

Sub

stra

te c

on

vers

ion

[%

]

Reaction time [h]

Uracil/(dUrd2‘F + Uracil) × 100 %

dAdo2‘F/(Adenine + dAdo2‘F) × 100 %

TtPyNP + GtPNP 70 °C TtPyNP + ApMTAP 80 °C

Figure 50: Progress curves for the enzymatic synthesis of dAdo2′F mediated by TtPyNP + GtPNP at 70 °C or by TtPyNP + ApMTAP at 80 °C. Pentofuranosyl donor was dUrd2′F and acceptor was adenine.

The synthesis of dAdo2′F is not included in the patent, but the example of 2,6-diamino-9-(2-deoxy-2-

fluoro-β-D-arabinofuranosyl)-9H-purine likewise illustrates the synthetic challenges encountered

with EcPNP and EcTP as biocatalyst. The enzyme loading was extremely high, extending even the

load that was used by the same group for the synthesis of dAdo2′F (Tuttle et al. 1993). For example

2900 U ml-1 of EcPNP were used by Tuttle and Krenitsky for 2,6-diamino-9-(2-deoxy-2-fluoro-β-D-

arabinofuranosyl)-9H-purine synthesis, while in the equivalent reaction for dAdo2′F synthesis only

195 U ml-1of EcPNP were used. With the assumptions and calculations outlined in the previous

section, an enzyme loading of 2900 U ml-1 would correspond to more than 12.6 mg ml-1.

Furthermore, the enzymes were not used in soluble form but as immobilized catalysts – presumably

because the reaction was conducted at 50 °C in order to increase the reaction rates as described in

(Tuttle et al. 1993). Nevertheless, the reaction required a relatively long incubation time of 6 days.

From the amount of isolated dAdo2′F (0.2 g) one can calculate the isolated yield with respect to 1-(2-

deoxy-2-fluoro-β-D-arabinofuranosyl)thymine, which is approx. 62 %. In the present study the yield

of dAdo2′F from adenine was 24 % after only 1 day, whereby the loading with each enzyme was only

0.1 mg ml-1. Hence, it appears that the high thermal stability and the favourable substrate specificity

of TtPyNP and ApMTAP indeed translate into considerably improved biocatalytical efficiencies.

Yamada and co-workers have described the synthesis of dAdo2′F from chemically prepared 2-deoxy-2-

fluoroarabinofuranosyl-α-1-phosphate and adenine as substrates, whereby the purine base was

applied in excess in the reaction mixture (5 mM versus 2.4 mM). The formation of the gycosidic bond

was catalyzed by PNP from G. stearothermophilus that allowed the operation at 50 °C without the

need of enzyme stabilization measures. With this strategy the yield of dAdo2′F with respect to 2-

deoxy-2-fluoroarabinofuranosyl-α-1-phosphate was about 50 %, reached after only 2.5 days. Even

though this process seems significantly more efficient than the enzymatic transglycosylation

described by Tuttle et al. with E. coli enzymes, the disadvantage lies in the complexity of the chemical

synthesis of 2-deoxy-2-fluoroarabinofuranosyl-α-1-phosphate. In fact, the authors motivated their

work with the reason that the rate of the phosphorolysis reaction in the enzymatic transglycosylation

Enzymatic transglycosylations with thermostable NPs 107

is very low. In this respect, the enzymatic transglycosylation employing the thermostable biocatalysts

(ApMTAP and TtPyNP) described in the present work represents a promising alternative.

5.3. Synthesis of 2,6-dihalogenated purine nucleosides

After having investigated the synthesis of carbohydrate-modified adenosine analogues, the next aim

was to test the applicability of the thermostable NPs for the synthesis of base-modified purine

nucleosides, namely 2,6-dichloropurine and 6-chloro-2-fluoropurine ribosides and deoxyribosides,

respectively. Since these target compounds were hardly available, the suitability of the enzymes with

PNP activity was investigated directly in the synthetic reaction, employing pentose-1-phosphate and

the modified purine base. Pentose-1-phosphates were thereby provided by the PyNP-catalyzed

phosphorolysis of uridine or thymidine. The reactions scheme for the synthesis of 2,6-dihalogenated

purine ribosides is illustrated in Figure 51. The reaction scheme for the respective deoxyribosides is

basically the same, with the exception that thymidine serves as pentofuranosyl donor.

To our best knowledge, enzymatic transglycosylations affording the 2,6-dihalogenated ribosides and

deoxribosides that are subject of this section have not been previously disclosed in the scientific

literature.

O

HO OH

HO

NH

N

O

O

+ Pi+ Pi

Uracil

PyNPPNP/MTAP

NH

NH

O

O

O

HO OH

HO

OPO32-

O

HO OH

HO

N

NN

N

Cl

Cl

Uridine

2,6-Dichloropurine

N

NNH

N

Cl

Cl

N

NNH

N

Cl

F

6-Chloro-2-fluoropurine

O

HO OH

HO

N

NN

N

Cl

F

2,6-Dichloropurine riboside

6-Chloro-2-fluoropurine riboside

Figure 51: Reaction scheme for the enzymatic synthesis of 2,6-dihalogenated ribosides. Uridine serves as pentofuranosyl donor, while either 2,6-dichloropurine or 6-chloro-2-fluoropurine functions as pentofuranosyl acceptor. The enzymatic transfer reaction can be catalyzed by the concurrent activity of PyNP and PNP, or MTAP respectively.

108 Enzymatic transglycosylations with thermostable NPs

5.3.1. Synthesis of 2,6-dihalogenated purine ribosides

Different combinations of enzymes were tested for their ability to catalyze the enzymatic synthesis of

2,6-dichloropurine riboside and 6-chloro-2-fluoropurine riboside. Uridine (2 mM) was used as ribose-

1-phosphate donor; the purine bases (2,6-dichloropurine, 6-chloro-2-fluoropurine) were employed in

1 mM concentration. In order to avoid spontaneous reactions of the halogenated purine bases, the

reaction temperature was restricted to a maximum of 65 °C. For screening the efficiency of different

combinations of biocatalysts, the reaction mixtures were analyzed after an incubation time of

30 min.

Indeed with all combinations of enzymes tested new peaks emerged, indicating the formation of 2,6-

dichloropurine riboside (retention time 8.2 min) and 6-chloro-2-fluoropurine riboside (7.6 min).

Authentic samples of the products were not available, but product formation as shown in (Figure 52)

could be indirectly quantified from the decrease of the concentration of the heterocyclic base. In

control experiments employing the natural purine base adenine as pentofuranosyl acceptor,

adenosine was obtained in very high yield with respect to the base (approx. 90 %). If the

dihalogenated purine were applied as pentofuranosyl acceptor, the yield with respect to the base

was between 49 % and 65 % for all the different combinations of enzymes tested. The highest

substrate conversion was obtained in reactions involving GtPNP, but a striking difference between

the different enzymes with PNP activity was not disclosed. The formation of adenosine was also

efficient when a PyNP was employed as sole biocatalyst. At 65 °C for example the same fraction of

adenine molecules (90 %) were converted to adenosine after 30 min, whether or not a second

enzyme with PNP activity was added to the reaction mixture (Figure 52B). These results further

corroborate the PNP by-activity of the PyNPs as discussed in section 4.1.5. However, if the artificial

halogenated purine bases were employed as pentofuranosyl acceptors the product formation was

dramatically lower in reactions with only PyNP as biocatalyst compared to the 2-enzyme reactions

(Figure 52). The use of a PNP enzyme (DgPNP, GtPNP or ApMTAP) as sole catalyst did not afford any

substrate conversion.

The transglycosylation reactions employing GtPNP and GtPyNP at 65 °C and artificial purine bases as

pentofuranosyl acceptors were further investigated by monitoring the progress of the reactions over

time. In addition transglycosylations with TtPyNP and ApMTAP were monitored at 80 °C over a 2 h

period, since the stability test had meanwhile shown that the dihalogenated purine bases did not

react to side products under these conditions. The results show that an equilibrium of the reaction

was relatively rapidly established, within a 1 – 2 h period (Figure 53). Further incubation did not

increase the product concentration. In the reactions where 2,6-dichloropurine was used applied as

pentofuranosyl acceptor, about 56 mol % of the purine base molecules were converted to the

corresponding riboside. Apparently, the only difference between the reactions employing

GtPNP/GtPyNP at 65 °C and TtPyNP/ApMTAP at 80 °C is that the equilibrium is faster reached in the

latter case (Figure 53A,B). Basically the same phenomenon was observed with 6-chloro-2-

fluoropurine, although the final outcome was significantly better: 63 – 67 mol % of base molecules

were transformed into the respective ribonucleoside (Figure 53C,D).

Enzymatic transglycosylations with thermostable NPs 109

0

20

40

60

80

100

GtPyNP GtPyNP + DgPNP

GtPyNP GtPyNP + DgPNP

GtPyNP GtPyNP + DgPNP

Adenine 2,6DCP 6C2FP

Sub

stra

te c

on

vers

ion

[m

ol

%]

Uridine Phosphorolysis

Glycosylated base

Uracil/(Uridine+ Uracil) × 100 %

Product/(Purine base + Product) × 100 %

0

20

40

60

80

100

GtPyNP GtPyNP + GtPNP

GtPyNP + ApMTAP

GtPyNP GtPyNP + GtPNP

GtPyNP + ApMTAP

GtPyNP GtPyNP + GtPNP

GtPyNP + ApMTAP

Adenine 2,6DCP 6C2FP

Sub

stra

te c

on

vers

ion

[m

ol

%]

0

20

40

60

80

100

TtPyNP TtPyNP + GtPNP

TtPyNP TtPyNP + GtPNP

TtPyNP TtPyNP + GtPNP

Adenine 2,6DCP 6C2FP

Sub

stra

te c

on

vers

ion

[m

ol

%]

A)

B)

C)

Figure 52. Synthesis of 2,6-dihalogentated purine ribosides with (A) DgPNP and GtPyNP at 55 °C, (B) GtPNP or ApMTAP in combination with GtPyNP at 65 °C and (C) GtPNP and TtPyNP at 65 °C . Uridine (2 mM) served as pentofuranosyl donor; adenine, 2,6-dichloropurine (2,6DCP) and 6-chloro-2-fluoropurine (6C2FP) were investigated as acceptors (1 mM). Transglycosylations were performed in 2 mM sodium phosphate buffer, pH 6.5. Reactions were stopped after 30 min.

110 Enzymatic transglycosylations with thermostable NPs

56

0

20

40

60

80

0 1 2 3 4 5 6 7

Sub

stra

te c

on

vers

ion

[m

ol

%]

Reaction time [h]

GtPyNP + GtPNP 65°C

56

0

20

40

60

80

0 1 2

Sub

stra

te c

on

vers

ion

[m

ol

%]

Reaction time [h]

TtPyNP + ApMTAP 80°C

Uracil

26DCP riboside

Uracil/(Uridine+ Uracil) × 100 %

2,6DCPriboside/(2,6DCP + 2,6 DCPriboside) × 100 %

54

63

0

20

40

60

80

0 1 2

Sub

stra

te c

on

vers

ion

mo

l [%

]

Reaction time [h]

TtPyNP + ApMTAP 80°C

Uracil

6C2FP riboside

54

67

0

20

40

60

80

0 1 2 3 4 5 6 7

Sub

stra

te c

on

vers

ion

[m

ol

%]

Reaction time [h]

GtPyNP + GtPNP 65°C

Uracil/(Uridine+ Uracil) × 100 %

6C2FPriboside/(6C2FP + 6C2FPriboside) × 100 %

A) B)

C) D)

Figure 53: Synthesis of dihalogenated ribosides over time. Formation of 2,6-dichloropurine riboside employing GtPNP + GtPyNP at 65 °C (A) or TtPyNP + ApMTAP at 80 °C (B); formation of 6-chloro-2-fluoropurine riboside employing GtPNP + GtPyNP at 65 °C (C) or TtPyNP + ApMTAP at 80 °C (D). Uridine (2 mM) served as pentofuranosyl donor; adenine, 2,6-dichloropurine (2,6DCP) and 6-chloro-2-fluoropurine (6C2FP) were investigated as acceptors (1 mM). Transglycosylations were performed in 2 mM sodium phosphate buffer, pH 6.5.

5.3.2. Synthesis of 2,6-dihalogenated purine deoxyribosides

In analogy to the 2,6 dihalogenated ribosides the synthesis of the deoxyriboside congeners was

investigated. Hereby thymidine served as pentofuranosyl donor. Initial tests revealed that the

synthesized purine deoxyribosides were not stable under acidic conditions and hydrolyzed to

pentose and purine base moieties (data not shown). For these reasons stopping the reactions

through addition of TCA proved not to be appropriate. Instead samples were now 3-fold diluted with

ice-cold buffer. Two different enzyme combinations were tested: TtPyNP + GtPNP at 65 °C and

TtPyNP + ApMTAP at 80 °C and the reaction mixture was analyzed after 30 min. With both

combinations new peaks emerged in the HPLC chromatogram of respective reaction mixture

samples, suggesting the formation of 2,6-dichloropurine deoxyriboside (retention time 8.8 min) and

6-chloro-2-fluoropurine deoxyriboside (8.2 min). Authentic samples of the products were not

available; product formation was thus again just indirectly quantified from the decrease of the

concentration of the heterocyclic base.

Monitoring of the substrate conversion over time reveals an interesting development in all reaction

mixtures studied here (Figure 54): Instead of increasing or constant product concentration, a

Enzymatic transglycosylations with thermostable NPs 111

decrease of the product concentrations over time was observed. At the same time the ratio of

thymine/thymidine increased. These results indicate that either the product or the intermediate

product (deoxyribose-1-phosphate) is instable and possible hydrolyzed. If for example the phosphate

would be cleaved form the deoxyribose-1-phosphate, this intermediate product would not be

anymore substrate for both the PyNP and the PNP catalyzed reactions. As a consequence both

enzymes would irreversibly catalyze the phosphorolysis reactions. Presumably it is indeed the

instability of the pentose-1-phosphate that is responsible for the observed effect. This would also be

an explanation for a previously observed effect. In Figure 38 it is shown that almost 100 % thymidine

molecules were phosphorolyzed at 80 °C with TtPyNP, while at 60 °C or with GtPyNP as biocatalyst

apparently an equilibrium with about 75 % phosphorolyzed molecules was reached. Also in the

transglycosylation reactions investigated now, the (intermediate) product instability appears to be

more pronounced at higher temperature; in both cases the apparent product decay was significantly

faster at 80 °C than at 65 °C (Figure 54).

0

20

40

60

80

100

0 1 2

Sub

stra

te c

on

vers

ion

[m

ol

%]

Reaction time [h]

TtPyNP + ApMTAP 80°CThymidine

26DCPdeoxyriboside

0

20

40

60

80

100

0 2 4 6

Sub

stra

te c

on

vers

ion

[m

ol

%]

Reaction time [h]

GtPyNP + GtPNP 65°C Thymine/(Thymidine+ Thymine) × 100 %

Product/(2,6DCP + Product) × 100 %

0

20

40

60

80

100

0 1 2

Sub

stra

te c

on

vers

ion

[m

ol

%]

Reaction time [h]

TtPyNP + ApMTAP 80°CThymidine

6C2FP deoxyriboside

0

20

40

60

80

100

0 2 4 6

Sub

stra

te c

on

vers

ion

mo

l [%

]

Reaction time [h]

GtPyNP+GtPNP 65°CThymine/(Thymidine+ Thymine) × 100 %

Product/(6C2FP + Product) × 100 %

A) B)

C) D)

Figure 54: Synthesis of dihalogenated deoxyribosides. Formation of 2,6-dichloropurine deoxyriboside employing GtPyNP + GtPNP at 65 °C (A) or TtPyNP + ApMTAP at 80 °C (B) and formation of 6-chloro-2-fluoropurine deoxyriboside employing GtPyNP + GtPNP at 65 °C (C) or TtPyNP + ApMTAP at 80 °C (D). Thymidine (2 mM) served as pentofuranosyl donor and 2,6-dichloropurine (2,6DCP) as acceptors (1 mM). Transglycosylations were performed in 2 mM sodium phosphate buffer, pH 6.5. Reactions were stopped after defined time intervals through the addition of 2 volumes of ice-cold buffer to 1 volume reaction mixture and further subsequent cooling.

112 Enzymatic transglycosylations with thermostable NPs

Future starting points for increasing the yield of 2,6-dihalogenated purine deoxyribosides

Due to the apparent lability of the intermediate product (2-deoxyribofuranosyl -α-1-phosphate) a

high temperature approach appears not to be suitable for the synthesis of the purine deoxyribosides

here under investigation and the reactions should be repeated employing DgPNP/GtPyNP at 55 °C.

For further optimization other factors that might have a positive effect on the stability of the

(intermediate) product might helpful to study (buffer composition, pH). Another option would be to

precisely determine when it would be best to stop the reaction. In the examples shown with GtPyNP

and GtPNP at 65 °C, it would thus appear appealing to stop the reaction after only 30 min, which

would result in approximately 70 % yield of nucleoside from the respective base. The use of

immobilized enzyme preparations could facilitate such an approach. A simple technical solution

could thus represent the batch-wise operation in a stirred tank reactor, in which the immobilized

enzymes are retained after product recovery, e.g. by a bottom stainless steel screen (Illanes and

Altamirano 2008). By the instant removal of the biocatalysts the presumably labile intermediate

product (deoxyribose-α-1-phosphate) would no longer be formed and the product concentration

should not decrease. A further advantage of this strategy would be that the biocatalyst could be

repeatedly used and product contamination with the biocatalysts would be avoided. It is also

conceivable to use immobilized biocatalysts in a continuous approach, for example in a stirred tank

reactor or in a packed column through which the substrate stream passes (Illanes and Altamirano

2008). However, continuous processes are usually selected for rather large scale productions and

hence the target production scale should be determined and considered for choosing the most

adequate reactor configuration.

5.4. Enzymatic transglycosylations– summary and conclusions

The results presented in this chapter have shown that the thermostable NPs were successfully

employed as biocatalysts for the enzymatic synthesis of both i) sugar modified purine nucleosides

(dAdo2′F, dAdo2′F)and for the synthesis of ii) purine nucleosides with modified purine base

(2,6-dihalogenated ribosides or deoxyribosides).

5.4.1. 2′-Fluorinated purine nucleosides

2′-Fluorinated adenine nucleosides (dAdo2′F, dAdo2′F) were synthesized via enzymatic

transglycosylation reactions employing 2′-fluorinated deoxyuridines (dUrd2′F, dUrd2′F) as

pentofuranosyl donors and adenine as pentofuranosyl acceptor. Different combinations of the

thermostable NPs that have been generated were investigated towards their suitability to catalyze

the reactions.

Synthesis of dAdo2′F

The combination of TtPyNP and DgPNP employed at 55 °C was found as most suitable fordAdo2′F

synthesis. Apart from the final yield, the increasing side product formation from dUrd2′F at higher

temperature was considered for the evaluation of enzyme combinations. With TtPyNP and DgPNP at

Enzymatic transglycosylations with thermostable NPs 113

55 °C, the final yield of dAdo2′F from adenine was 23 % after 1 day. The fact that the product

concentration increased in quasi linear relation to time indicates that the final outcome was limited

by the reaction rates of the enzymes. Therefore, higher enzyme loadings and prolonged reaction

times would presumably easily lead to higher yields. Moreover the ratio of the substrates and

enzymes in the reaction mixture could be optimized in future experiments.

The presented results should therefore be regarded as preliminary results, indicating the feasibility of

the approach to use thermostable enzymes for dAdo2′F synthesis. However, already at this stage the

results show that the efficiency of the TtPyNP/DgPNP enzyme preparation in the synthesis of dAdo2′F

is presumably higher than the efficiency of Lactobacillus reuteri NdRT that was recently described as

biocatalyst for the same reaction (Fernandez-Lucas et al. 2010).

Comparing the results to the data reported by Tuttle et al. (82 % yield with respect to dUrd2′F after 17

days), who have accomplished the transglycosylation reaction employing EcPNP and EcTP at

37 °C(Tuttle et al. 1993),it is difficult owing to the different reaction conditions that were used (molar

ratio of the substrates, different enzyme loading). In fact, the final yield reported by Tuttle is higher

than that obtained in the present study. However, taking into account that reaction time and enzyme

loading were both significantly higher, we can speculate that the combination of TtPyNP and DgPNP

at 55 °C has the potential to catalyze the same reaction with significantly higher efficiency. Future

investigations should include experiments with molar substrate ratios as used by Tuttle and co-

workers, higher enzyme loadings and longer reaction times. The results will then show whether the

use of TtPyNP/DgPNP is advantageous over the use of the E. coli enzymes.

Synthesis of dAdo2′F

The best conditions for dAdo2′F synthesis were found with TtPyNP/ApMTAP at 80 °C, which lead to

24 % yield after 1 day. Again the product concentration increased quasi linear in relation to time.

Hence, same measures as proposed for dAdo2′F synthesis (higher enzyme loading, prolonged reaction

times) would presumably also lead here to and improved final outcome.

However, already the preliminary results presented here indicate that the thermostable enzyme

preparation (TtPyNP+ ApMTAP) applied, seems to be more efficient than the combination of EcPNP

and EcTP used by Tuttle and Krenitsky in 1992 for the synthesis of 9-(2-deoxy-2-fluoro-β-D-

arabinofuranosyl)purines via enzymatic transglycosylation.

On the other hand, the advantage compared to the method developed by Yamada et al. (Yamada et

al. 2009), who have used GsPNP to catalyze the synthesis of dAdo2′F from 2-deoxy-2-

fluoroarabinofuranosyl-α-1-phosphate and adenine, is that the substrate used in the present study

(dUrd2′F) is presumably easier chemically prepared than 2-deoxy-2-fluoroarabinofuranosyl-α-1-

phosphate.

5.4.2. 2,6-Dihalogenated purine nucleosides

2,6-Dihalogenated purine nucleosides were synthesized via enzymatic transglycosylation reactions

employing either thymidine or uridine as pentofuranosyl donor and 2,6-dichloropurine or 6-chloro-2-

114 Enzymatic transglycosylations with thermostable NPs

fluoropurine as pentofuranosyl acceptor. Note that authentic samples of the products were not

available as reference and the identity of the products should hence be confirmed by adequate

analytical investigations in future experiments.

2,6-Dihalogenated purine ribosides

Different combinations of the thermostable NPs were investigated towards their suitability to

catalyze the synthesis of 2,6-dichloropurine riboside and 6-chloro-2-fluoropurine riboside via

enzymatic transglycosylations. The results indicate that the situation is very different from that

observed concerning the synthesis of 2′-fluorinated purine nucleosides. Obviously, the equilibrium of

the transglycosylation reaction was rather rapidly reached, with the consequence that the final

outcome evaluated within a 6 h period seems not to depend significantly on the choice of

temperatures and enzyme combination. Instead, rather the ratio of product and substrate

concentrations at equilibrium conditions seems to limit the final yield (56 % 2,6-DCPriboside,

63 - 67 % 6C2FP riboside with respect to purine base). Future optimization studies should therefore

target the equilibrium conditions rather than the enzyme loading and extended reaction times as

proposed for the synthesis of 2′-fluorinated nucleosides. Possible starting points include the pH and

the concentration of the phosphate buffer, as well as the ratio of pentofuranosyl donor and

acceptor. Furthermore, other ribofuranosyl donors that are essentially irreversibly phosphorolyzed

could be considered in order to favourably shift the equilibrium conditions.

2,6-Dihalogenated purine deoxyribosides

Within the experiments aiming at the synthesis of 2,6-dihalogenated purine deoxyribosides a totally

different obstacle was encountered. Apparently, the intermediate compound 2-deoxyribofuranosyl -

α-1-phosphate was not stable, leading to the exclusive formation of thymine and dihalogenated

purine after prolonged reaction times. Further optimizations should therefore aim on conditions that

are favourable for the stability of 2-deoxyribofuranosyl-α-1-phosphate, for example reduced reaction

temperatures, or different buffer compositions. Additionally, reactions should not proceed to long

but rather stopped after a defined period. This could be realized by making use of immobilized

enzymes. An advantage of this approach would be that the recovered biocatalysts could be re-used

and that a number of reactor configurations and operation modes would be conceivable.

6. Final conclusions

Key objectives of the present work have been accomplished and concern: i) the recombinant

expression of thermostable nucleoside phosphorylases (NPs) in E. coli, ii) the characterization of the

generated NPs with respect to potential biocatalytical applications and finally iii) the synthesis of

modified purine nucleosides (2′-fluorinated and 2,6-dihalogenated purine nucleosides) that are

difficult to access by chemical methods. In the following sections, the main conclusions that can be

drawn from the course of this study will be listed.

Recombinant expression of thermostable NPs in E. coli

Six NPs from thermophilic microorganisms have been selected for the recombinant

expression in E. coli: purine nucleoside phosphorylase (PNP) from Deinococcus geothermalis

and from Geobacillus thermoglucosidasius, 5'-Methylthioadenosine phosphorylase (MTAP)

and uridine phosphorylase (UP) from Aeropyrum pernix, and pyrimidine nucleoside

phosphorylase (PyNP) from Geobacillus thermoglucosidasius and Thermus thermophilus.

Except ApUP all enzymes have been successfully overexpressed in biologically active form in

E. coli. The expression vector confers a hexahistidine tag that is fused to the N-terminus and

contains an IPTG inducible lac promoter.

Soluble expression levels were very different, strongly depending on the target enzyme. The

expression of GtPNP and GtPyNP was very robust, resulting in high yields of soluble

recombinant enzyme and was relatively independent of the specific expression conditions

chosen. By contrast, expression of DgPNP and ApMTAP was slightly less efficient and more

prone to aggregation. Here, the expression and/or final cell density were optimized by fine-

tuning the inducer concentration and controlling the growth rate by employing medium with

enzyme-based glucose delivery. Expression of TtPyNP was more challenging and could not be

considerably improved. Most reproducible conditions were found by expression with

enzyme-based glucose delivery.

Initial expression studies have demonstrated the sensitivity of DgPNP towards N-terminal

fusion tags. We reasoned that excessive additional amino acids conferred by recombinational

cloning into a “destination” vector rendered PNP prone to aggregation. To avoid similar

problems with other NPs the respective expression construct was not further used and genes

were rather cloned by the conventional strategy involving restriction and ligation.

When native sequences of DgPNP and ApMTAP without any fusion tag were expressed, the

formation of stable secondary 5′ mRNA structures was found to impair the efficient

expression. The strategy to overcome the problem studied here, involved sequence

optimization and high-temperature cultivation (up to 42 °C) of the E. coli expression strains.

It seems likely that this strategy would also translate into better yields in the expression of

other thermostable proteins expressed without artificial N-terminal tag.

116 References

Characterization of thermostable NPs

Thermal properties of NPs varied according to the source microorganism and type of

enzyme. The following temperature optima were determined: 55 °C (DgPNP), 60 °C (GtPyNP)

and 70 °C (GtPNP). In case of TtPyNP and ApMTAP the reaction rate increased up to the

highest temperature tested. The degree of thermal stability correlated with the temperature

optima. DgPNP represents the least thermostable enzyme (t1/2 = 1.7 h at 60 °C) whereas

ApMTAP is the most thermostable NP here under investigation (t1/2 > 27 h at 90 °C).

As it is expected from the gene annotations the PyNPs readily catalyze the phosphorolytic

cleavage of uridine and thymidine, whereas the PNPs recognize both inosine and adenosine

as substrate, which is typical for bacterial type PNPs. The PNP substrate specificities are also

in agreement with the phylogenetic analysis that was based on amino acid alignments to

studied PNPs. ApMTAP is phylogenetically closely related to MTAPI from

Sulfolobus solfataricus and shows likewise similar substrate activities (accepting both inosine

and adenosine).

Unexpectedly, purine nucleosides (inosine, adenosine) were also recognized as substrates by

the PyNPs, even there was a clear preference for uridine and thymidine. This “PNP” activity

of enzymes with the NP-II fold is best to our knowledge first time described.

Cytidine was weakly recognized as substrate by PNPs and ApMTAP, whereas the cytidine

phosphorolysis rate determined for TtPyNP was significantly lower. For GtPyNP no substrate

activity of cytidine was seen.

The fluorination of the 2′-position of uridine leads to a dramatic decrease of substrate activity

for PyNPs, but is significantly better tolerated by TtPyNP than by GtPyNP. Fluorination of the

arabino position (dUrd2′F) is more critical, than the fluorination of the ribo position (dUrd2′F).

In fact, 2′-arabino fluorinated uridine is not recognized at all by GtPyNP as substrate.

In analogy, the phosphorolysis of adenosine analogues, fluorinated on the 2′ position, by NPs

with PNP activity (DgPNP, GtPNP, ApMTAP) was investigated. The data show that the highest

absolute phosphorolysis rates for both compounds (dAdo2′F and dAdo2′F) were obtained with

DgPNP. By contrast, GtPNP, that actually appears to be the most efficient catalyst for natural

purine nucleosides (inosine, adenosine) shows less activity than DgPNP for dAdo2′F and is the

weakest catalyst for dAdo2′F.

Kinetic parameters of PyNPs were determined for the natural substrates uridine and

thymidine at 60 °C (GtPyNP) and 80 °C (TtPyNP). In contrast to TtPyNP, GtPyNP is

characterized by extremely low substrate affinities (high Km values) towards both natural

substrates. The turnover numbers (kcat) of GtPyNP and TtPyNP are in similar range, with

uridine as substrate. By contrast, the turnover numbers for thymidine differ significantly, in

favour of thymidine phosphorolysis by TtPyNP. The kcat value of TtPyNP for thymidine is also

unusually high in comparison to EcTP, which demonstrates the potential for biocatalytic

applications.

Enzymatic transglycosylations with thermostable NPs 117

Enzymatic transglycosylations with thermostable NPs

The generated thermostable NPs were successfully applied in transglycosylation reactions

affording purine nucleoside analogues.

dAdo2′F was synthesized from dUrd2′F (pentofuranosyl donor) and adenine (pentofuranosyl

acceptor). The combination of TtPyNP and DgPNP applied at 55 °C was found to be the best

strategy for dAdo2′F synthesis. The operation at considerably higher temperatures proved to

be not feasible in reactions employing dUrd2'Fas substrate since dUrd2'F is relative labile,

reacting to unwanted side products at high temperature.

The synthesis of dAdo2'F was most efficiently catalyzed by ApMTAP + TtPyNP at 80 °C (from

dUrd2'F as donor and adenine as pentofuranosyl acceptor.

The efficient synthesis of sugar-fluorinated nucleosides was generally limited by the reaction

rates of the enzymes. Therefore it is expected that higher enzyme loadings and prolonged

incubation times will lead to higher yield. However, already the preliminary results presented

here, indicate that the use of the thermostable enzyme preparations instead of

conventionally used E. coli enzymes permits the operation at higher temperature and

translates into increased efficiencies of the respective transglycosylation reactions.

Both, 2,6-dichloropurine and 6-chloro-2-fluoropurine are well accepted as substrates by NPs

with PNP activity here under investigation (DgPNP, GtPNP, ApMTAP). The respective

ribosides can hence be readily prepared by employing uridine as pentofuranosyl donor and a

combination of PyNP and PNP (ApMTAP). In contrast to the synthesis of sugar-fluorinated

purine nucleosides, the final outcome (56 % to 67 % yield with respect to the purine base) is

not limited by the reaction rate of the enzymes, but rather by the conditions of the

equilibrium, that is rapidly established.

The synthesis of 2,6-dihalogenated deoxyribosides from thymidine as pentofuranosyl donor

and dihalogenated purine bases as pentofuranosyl acceptor is likewise efficiently catalyzed

by the thermostable NPs investigated in the present study. However, prolonged reaction

times lead to the exclusive formation of thymine and dihalogenated purine. This

phenomenon is accelerated at higher temperature and is presumably due to the instability of

the intermediate product deoxyribofuranosyl -α-1-phosphate.

In control experiments transglycosylation reactions were also performed by employing either

PyNP or PNP (ApMTAP) as sole biocatalyst. With the exclusive use of PNPs or ApMTAP as

biocatalyst, no product formation was observed. By contrast, owing to the PNP side activity

of the PyNPs, natural purine nucleosides (inosine, adenosine) were efficiently synthesized, by

applying PyNP as sole biocatalyst. However, this catalytic activity of the PyNP was not of

practical use for the synthesis of purine nucleoside analogues, since substrate activities of

modified purine nucleosides were very poor for PyNPs.

118 References

Outlook

In summary, the data demonstrate the promising potential of the investigated thermostable NPs for

applications aiming at the synthesis of unnatural nucleosides from chemically modified precursors.

Future experiments may be envisioned to be devoted to i) the optimization of the synthetic reactions

presented in this work and to ii) the use of the repertoire of thermostable NPs for the disclosure of

novel reactions.

A number of variables appear to be promising to study in regard to the optimization of the

transglycosylation reactions. Examples include the variation of substrate ratios (pentofuranosyl

donor versus acceptor), the buffer composition (concentration of phosphate ions, pH) and the

enzyme loading ratios. In order to better understand the nature of the reactions, it might

furthermore be of advantage if the reactions would be mathematically modelled. Such an approach

could also pave the way for in silico optimizations. Since purine bases are in general only poorly

soluble in aqueous solutions, and thermostability is often correlated to higher stability of enzymes in

the presence of organic solvents, it would furthermore be appealing to investigate enzymatic

transglycosylations in organic solvents. Finally the use of immobilized enzyme preparations would be

advantageous in regard to industrial applications.

Eventually it would be interesting to expand the repertoire of nucleoside products that can be

synthesized through the use of the generated NPs. After having shown here that both sugar-modified

and base modified purine nucleosides can be synthesized it would be conceivable to investigate the

synthesis of purine nucleosides that are modified on both moieties at the same time. Furthermore,

extensive substrate activity studies with chemically prepared precursors could reveal which

substrates are readily recognized and clarify the scope of feasible future applications.

References

Adler, A., Forster, N., Homann, M., and Goringer, H.U., 2008. "Post-SELEX chemical optimization of a trypanosome-specific RNA aptamer". Comb Chem High Throughput Screen, 11(1), 16-23

Agrawal, S., 2010. "Remembering Paul C. Zamecnik, M.D., "Father of Antisense" (1912-2009)". Oligonucleotides, 20(1), 47-50

Akashi, H., 1994. "Synonymous Codon Usage in Drosophila-Melanogaster - Natural-Selection and Translational Accuracy". Genetics, 136(3), 927-935

Akiyama, S., Furukawa, T., Sumizawa, T., Takebayashi, Y., Nakajima, Y., Shimaoka, S., and Haraguchi, M., 2004. "The role of thymidine phosphorylase, an angiogenic enzyme, in tumor progression". Cancer Sci, 95(11), 851-857

Alexeev, C.S., Panova, N.G., Polyakov, K.M., and Mikhailov, S.N., 2010. "Substrate specificity of E. coli uridine phosphorylase". Abstracts of XIX International Round Table on Nucleosides, Nucleotides and Nucleic Acids, IRT 2010(Lyon 2010), p. 94

Allerson, C.R., Sioufi, N., Jarres, R., Prakash, T.P., Naik, N., Berdeja, A., Wanders, L., Griffey, R.H., Swayze, E.E., and Bhat, B., 2005. "Fully 2'-modified oligonucleotide duplexes with improved in vitro potency and stability compared to unmodified small interfering RNA". J Med Chem, 48(4), 901-904

Almendros, M., Gago, J.V.S., and Carlos, J.B., 2009. "Thermus thermophilus Strains Active in Purine Nucleoside Synthesis". Molecules, 14(3), 1279-1287

Altschul, S.F., Gish, W., Miller, W., Myers, E.W., and Lipman, D.J., 1990. "Basic local alignment search tool". J Mol Biol, 215(3), 403-10

Angelov, A., Mientus, M., Liebl, S., and Liebl, W., 2009. "A two-host fosmid system for functional screening of (meta)genomic libraries from extreme thermophiles". Syst Appl Microbiol, 32(3), 177-85

Appleby, T.C., Mathews, I.I., Porcelli, M., Cacciapuoti, G., and Ealick, S.E., 2001. "Three-dimensional structure of a hyperthermophilic 5'-deoxy-5'-methylthioadenosine phosphorylase from Sulfolobus solfataricus". J Biol Chem, 276(42), 39232-39242

Baneyx, F., 1999. "Recombinant protein expression in Escherichia coli". Curr Opin Biotechnol, 10(5), 411-421

Baneyx, F. and Mujacic, M., 2004. "Recombinant protein folding and misfolding in Escherichia coli". Nat Biotechnol, 22(11), 1399-1408

Barai, V.N., Zinchenko, A.I., Eroshevskaya, L.A., Kalinichenko, E.N., Kulak, T.I., and Mikhailopulo, I.A., 2002. "A universal biocatalyst for the preparation of base- and sugar-modified nucleosides via an enzymatic transglycosylation". Helv Chim Acta, 85(7), 1901-1908

Bauta, W.E., Schulmeier, B.E., Burke, B., Puente, J.F., Cantrell, W.R., Lovett, D., Goebel, J., Anderson, B., Ionescu, D., and Guo, R.C., 2004. "A new process for antineoplastic agent clofarabine". Org Process Res Dev, 8(6), 889-896

Beeby, M., O'Connor, B.D., Ryttersgaard, C., Boutz, D.R., Perry, L.J., and Yeates, T.O., 2005. "The genomics of disulfide bonding and protein stabilization in thermophiles". PLoS Biol, 3(9), 1549-1558

Bellezza, I., Tucci, A., and Minelli, A., 2008. "2-Chloroadenosine and Human Prostate Cancer Cells". Anti-Cancer Agents Med Chem, 8(7), 783-789

Bennett, E.M., Li, C., Allan, P.W., Parker, W.B., and Ealick, S.E., 2003. "Structural basis for substrate specificity of Escherichia coli purine nucleoside phosphorylase". J Biol Chem, 278(47), 47110-47118

Berdis, A.J., 2008. "DNA polymerases as therapeutic targets". Biochemistry-Us, 47(32), 8253-8260 Bernstein, F.C., Koetzle, T.F., Williams, G.J., Meyer, E.F., Jr., Brice, M.D., Rodgers, J.R., Kennard, O.,

Shimanouchi, T., and Tasumi, M., 1977. "The Protein Data Bank: a computer-based archival file for macromolecular structures". J Mol Biol, 112(3), 535-542

120 References

Berrow, N.S., Bussow, K., Coutard, B., Diprose, J., Ekberg, M., Folkers, G.E., Levy, N., Lieu, V., Owens, R.J., Peleg, Y., Pinaglia, C., Quevillon-Cheruel, S., Salim, L., Scheich, C., Vincentelli, R., and Busso, D., 2006. "Recombinant protein expression and solubility screening in Escherichia coli: a comparative study". Acta Crystallogr D, 62(10), 1218-1226

Bonate, P.L., Arthaud, L., Cantrell, W.R., Jr., Stephenson, K., Secrist, J.A., 3rd, and Weitman, S., 2006. "Discovery and development of clofarabine: a nucleoside analogue for treating cancer". Nat Rev Drug Discov, 5(10), 855-863

Bordoli, L., Kiefer, F., Arnold, K., Benkert, P., Battey, J., and Schwede, T., 2009. "Protein structure homology modeling using SWISS-MODEL workspace". Nat Protoc, 4(1), 1-13

Brady, D., Jordaan, J., Simpson, C., Chetty, A., Arumugam, C., and Moolman, F.S., 2008. "Spherezymes: a novel structured self-immobilisation enzyme technology". BMC Biotechnol, 8(8),

Buckoreelall, K., Wilson, L., and Parker, W.B., 2011. "Identification and Characterization of Two Adenosine Phosphorylase Activities in Mycobacterium smegmatis". J Bacteriol, 193(20), 5668-5674

Bzowska, A., Kulikowska, E., and Shugar, D., 2000. "Purine nucleoside phosphorylases: properties, functions, and clinical aspects". Pharmacol Therapeut, 88(3), 349-425

Cacciapuoti, G., Porcelli, M., Bertoldo, C., Derosa, M., and Zappia, V., 1994. "Purification and Characterization of Extremely Thermophilic and Thermostable 5'-Methylthioadenosine Phosphorylase from the Archaeon Sulfolobus-Solfataricus - Purine Nucleoside Phosphorylase-Activity and Evidence for Intersubunit Disulfide Bonds". J Biol Chem, 269(40), 24762-24769

Cacciapuoti, G., Fusco, S., Caiazzo, N., Zappia, V., and Porcelli, M., 1999. "Heterologous expression of 5'-methylthioadenosine phosphorylase from the archaeon Sulfolobus solfataricus: characterization of the recombinant protein and involvement of disulfide bonds in thermophilicity and thermostability". Protein Expr Purif, 16(1), 125-135

Cacciapuoti, G., Bertoldo, C., Brio, A., Zappia, V., and Porcelli, M., 2003. "Purification and characterization of 5'-methylthioadenosine phosphorylase from the hyperthermophilic archaeon Pyrococcus furiosus - Substrate specificity and primary structure analysis". Extremophiles, 7(2), 159-168

Cacciapuoti, G., Moretti, M.A., Forte, S., Brio, A., Camardella, L., Zappia, V., and Porcelli, M., 2004. "Methylthioadenosine phosphorylase from the archaeon Pyrococcus furiosus - Mechanism of the reaction and assignment of disulfide bonds". Eur J Biochem, 271(23-24), 4834-4844

Cacciapuoti, G., Forte, S., Moretti, M.A., Brio, A., Zappia, V., and Porcelli, M., 2005. "A novel hyperthermostable 5'-deoxy-5'-methylthioadenosine phosphorylase from the archaeon Sulfolobus solfataricus". FEBS J, 272(8), 1886-1899

Cacciapuoti, G., Gorassini, S., Mazzeo, M.F., Siciliano, R.A., Carbone, V., Zappia, V., and Porcelli, M., 2007. "Biochemical and structural characterization of mammalian-like purine nucleoside phosphorylase from the Archaeon Pyrococcus furiosus". FEBS J, 274(10), 2482-2495

Cacciapuoti, G., Peluso, I., Fuccio, F., and Porcelli, M., 2009. "Purine nucleoside phosphorylases from hyperthermophilic Archaea require a CXC motif for stability and folding". FEBS J, 276(20), 5799-5805

Cacciapuoti, G., Marabotti, A., Fuccio, F., and Porcelli, M., 2011. "Unraveling the structural and functional differences between purine nucleoside phosphorylase and 5'-deoxy-5'-methylthioadenosine phosphorylase from the archaeon Pyrococcus furiosus". Bba-Proteins Proteom, 1814(10), 1358-1366

Cappellacci, L., Petrelli, R., Franchetti, P., Vita, P., Kusumanchi, P., Kumar, M., Jayaram, H.N., Zhou, B., Yen, Y., and Grifantini, M., 2011. "Synthesis and biological activity of novel N6-substituted and 2,N6-disubstituted adenine ribo- and 3'-C-methyl-ribonucleosides as antitumor agents". Eur J Med Chem, 46(5), 1499-1504

Care, S., Bignon, C., Pelissier, M.C., Blanc, E., Canard, B., and Coutard, B., 2008. "The translation of recombinant proteins in E. coli can be improved by in silico generating and screening random

References 121

libraries of a -70/+96 mRNA region with respect to the translation initiation codon". Nucleic Acids Res, 36(1), 1-6

Carson, D.A., Wasson, D.B., Kaye, J., Ullman, B., Martin, D.W., Jr., Robins, R.K., and Montgomery, J.A., 1980. "Deoxycytidine kinase-mediated toxicity of deoxyadenosine analogs toward malignant human lymphoblasts in vitro and toward murine L1210 leukemia in vivo". P Natl Acad Sci USA, 77(11), 6865-6869

Carson, D.A. and Wasson, D.B., 1988. "Synthesis of 2',3'-Dideoxynucleosides by Enzymatic Trans-Glycosylation". Biochem Bioph Res Co, 155(2), 829-834

Cassera, M.B., Hazleton, K.Z., Merino, E.F., Obaldia, N., 3rd, Ho, M.C., Murkin, A.S., DePinto, R., Gutierrez, J.A., Almo, S.C., Evans, G.B., Babu, Y.S., and Schramm, V.L., 2011. "Plasmodium falciparum parasites are killed by a transition state analogue of purine nucleoside phosphorylase in a primate animal model". PLoS One, 6(11), e26916

Cèbe, R. and Geiser, M., 2006. "Rapid and easy thermodynamic optimization of the 5'-end of mRNA dramatically increases the level of wild type protein expression in Escherichia coli". Protein Expres Purif, 45(2), 374-380

Chen, J., Acton, T.B., Basu, S.K., Montelione, G.T., and Inouye, M., 2002. "Enhancement of the solubility of proteins overexpressed in Escherichia coli by heat shock". J Mol Microbiol Biotechnol, 4(6), 519-524

Cieslinski, H., Dlugolecka, A., Kur, J., and Turkiewicz, M., 2009. "An MTA phosphorylase gene discovered in the metagenomic library derived from Antarctic top soil during screening for lipolytic active clones confers strong pink fluorescence in the presence of rhodamine B". FEMS Microbiol Lett, 299(2), 232-240

Codington, J.F., Fox, J.J., and Doerr, I.L., 1964. "Nucleosides. XVIII. Synthesis of 2'-Fluorothymidine, 2'-Fluorodeoxyuridine, and Other 2'-Halogeno-2'-Deoxy Nucleosides". J Org Chem, 29(3), 558-564

Condezo, L.A., Fernández-Lucas, J., García-Burgos, C.A., Alcántara, A.R., and Sinisterra, J.V., "Enzymatic Synthesis of Modified Nucleosides", in "Biocatalysts in the Pharmaceutical and Biotechnology Industries", R.N. Patel, Editor. 2006, Crc Press, Boca Raton. p. 403-426.

Copeland, R.A., "Enzymes: A Practical Introduction to Structure, Mechanism, and Data Analysis". Second ed. 2000, New York: Wiley-VCH.

Cowan, D.A., Arslanoglu, A., Burton, S.G., Baker, G.C., Cameron, R.A., Smith, J.J., and Meyer, Q., 2004. "Metagenomics, gene discovery and the ideal biocatalyst". Biochem Soc T, 32(Pt 2), 298-302

Daddona, P.E., Wiesmann, W.P., Milhouse, W., Chern, J.W., Townsend, L.B., Hershfield, M.S., and Webster, H.K., 1986. "Expression of Human Malaria Parasite Purine Nucleoside Phosphorylase in Host Enzyme-Deficient Erythrocyte Culture - Enzyme Characterization and Identification of Novel Inhibitors". J Biol Chem, 261(25), 1667-1673

Damha, M.J., Wilds, C.J., Noronha, A., Brukner, I., Borkow, G., Arion, D., and Parniak, M.A., 1998. "Hybrids of RNA and arabinonucleic acids (ANA and 2'F-ANA) are substrates of ribonuclease H". J Am Chem Soc, 120(49), 12976-12977

De Clercq, E., 2004. "Antiviral drugs in current clinical use". J Clin Virol, 30(2), 115-133 de Marco, A., 2009. "Strategies for successful recombinant expression of disulfide bond-dependent

proteins in Escherichia coli". Microb Cell Fact, 8(26), de Smit, M.H. and van Duin, J., 1990. "Secondary structure of the ribosome binding site determines

translational efficiency: A quantitative analysis". P Natl Acad Sci 87(19), 7668-7672 de Smit, M.H. and van Duin, J., 1994. "Control of translation by mRNA secondary structure in

Escherichia coli. A quantitative analysis of literature data". J Mol Biol, 244(2), 144-150 Deleavey, G.F., Watts, J.K., Alain, T., Robert, F., Kalota, A., Aishwarya, V., Pelletier, J., Gewirtz, A.M.,

Sonenberg, N., and Damha, M.J., 2010. "Synergistic effects between analogs of DNA and RNA improve the potency of siRNA-mediated gene silencing". Nucleic Acids Res, 38(13), 4547-4557

Dessanti, P., Zhang, Y., Allegrini, S., Tozzi, M.G., Sgarrella, F., and Ealick, S.E., 2012. "Structural basis of the substrate specificity of Bacillus cereus adenosine phosphorylase". Acta Crystallogr D, 68(Pt 3), 239-248

122 References

Dobak, I., Ostafin, M., Poleshchuk, O.K., Milecki, J., and Nogaj, B., 2008. "Reactivity of 2,6-Dichloropurine Ribonucleoside studied by 35Cl NQR Spectroscopy". Appl Magn Reson 34(1-2), 47-53

Dong, H.J., Nilsson, L., and Kurland, C.G., 1996. "Co-variation of tRNA abundance and codon usage in Escherichia coli at different growth rates". J Mol Biol, 260(5), 649-663

Donovan, R.S., Robinson, C.W., and Glick, B.R., 1996. "Review: Optimizing inducer and culture conditions for expression of foreign proteins under the control of the lac promoter". J Ind Microbiol, 16(3), 145-154

Esipov, R.S., Gurevich, A.I., Chuvikovsky, D.V., Chupova, L.A., Muravyova, T.I., and Miroshnikov, A.I., 2002. "Overexpression of Escherichia coli genes encoding nucleoside phosphorylases in the pET/Bl21 (DE3) system yields active recombinant enzymes". Protein Expres Purif, 24(1), 56-60

Esposito, D. and Chatterjee, D.K., 2006. "Enhancement of soluble protein expression through the use of fusion tags". Curr Opin Biotech, 17(4), 353-358

Fariselli, P. and Casadio, R., 1999. "A neural network based predictor of residue contacts in proteins". Protein Eng, 12(1), 15-21

Fernandez-Lucas, J., Acebal, C., Sinisterra, J.V., Arroyo, M., and de la Mata, I., 2010. "Lactobacillus reuteri 2'-Deoxyribosyltransferase, a Novel Biocatalyst for Tailoring of Nucleosides". Appl Environ Microb, 76(5), 1462-1470

Ferreira, A.C., Nobre, M.F., Rainey, F.A., Silva, M.T., Wait, R., Burghardt, J., Chung, A.P., and da Costa, M.S., 1997. "Deinococcus geothermalis sp. nov. and Deinococcus murrayi sp. nov., two extremely radiation-resistant and slightly thermophilic species from hot springs". Int J Syst Bacteriol, 47(4), 939-947

Fire, A., Xu, S.Q., Montgomery, M.K., Kostas, S.A., Driver, S.E., and Mello, C.C., 1998. "Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans". Nature, 391(6669), 806-811

Fiser, A., Cserzo, M., Tudos, E., and Simon, I., 1992. "Different Sequence Environments of Cysteines and Half Cystines in Proteins Application to Predict Disulfide Forming Residues". FEBS Lett, 302(2), 117-120

Flexner, C., 2007. "HIV drug development: the next 25 years". Nat Rev Drug Discov, 6(12), 959-966 Furman, P.A., Fyfe, J.A., Stclair, M.H., Weinhold, K., Rideout, J.L., Freeman, G.A., Lehrman, S.N.,

Bolognesi, D.P., Broder, S., Mitsuya, H., and Barry, D.W., 1986. "Phosphorylation of 3'-Azido-3'-Deoxythymidine and Selective Interaction of the 5'-Triphosphate with Human-Immunodeficiency-Virus Reverse-Transcriptase". P Natl Acad Sci USA, 83(21), 8333-8337

Furukawa, T., Yoshimura, A., Sumizawa, T., Haraguchi, M., Akiyama, S., Fukui, K., Ishizawa, M., and Yamada, Y., 1992. "Angiogenic Factor". Nature, 356(6371), 668-668

Gao, X.F., Huang, X.R., and Sun, C.C., 2006. "Role of each residue in catalysis in the active site of pyrimidine nucleoside phosphorylase from Bacillus subtilis: A hybrid QM/MM study". J Struct Biol, 154(1), 20-26

Gerdes, K., 1988. "The ParB (hok sok) Locus of Plasmid R1: a General Purpose Plasmid Stabilization System". Bio-Technol, 6(12), 1402-1405

Giblett, E.R., Ammann, A.J., Sandman, R., Wara, D.W., and Diamond, L.K., 1975. "Nucleoside-Phosphorylase Deficiency in a Child with Severely Defective T-Cell Immunity and Normal B-Cell Immunity". Lancet, 1(7914), 1010-1013

Gordon, G.E.R., Visser, D.F., Brady, D., Raseroka, N., and Bode, M.L., 2011. "Defining a process operating window for the synthesis of 5-methyluridine by transglycosylation of guanosine and thymine". J Biotechnol, 151(1), 108-113

Gragoudas, E.S., Adamis, A.P., Cunningham, E.T., Feinsod, M., Guyer, D.R., and Neova, V.I.S.O., 2004. "Pegaptanib for neovascular age-related macular degeneration". New Engl J Med, 351(27), 2805-2816

Green, E.A., Rosenstein, R.D., Shiono, R., Abraham, D.J., Trus, B.L., and Marsh, R.E., 1975. "Crystal-Structure of Uridine". Acta Crystallogr B, 31(Jan15), 102-107

Grillone, L.R. and Lanz, R., 2001. "Fomivirsen". Drugs Today, 37(4), 245-255

References 123

Griswold, K.E., Mahmood, N.A., Iverson, B.L., and Georgiou, G., 2003. "Effects of codon usage versus putative 5'-mRNA structure on the expression of Fusarium solani cutinase in the Escherichia coli cytoplasm". Protein Expr Purif, 27(1), 134-142

Gryaznov, S.M., Lloyd, D.H., Chen, J.K., Schulz, R.G., Dedionisio, L.A., Ratmeyer, L., and Wilson, W.D., 1995. "Oligonucleotide N3'-->P5' Phosphoramidates". P Natl Acad Sci USA, 92(13), 5798-5802

Gryaznov, S.M., 1999. "Oligonucleotide N3'-->P5' phosphoramidates as potential therapeutic agents". Biochim Biophys Acta, 1489(1), 131-140

Gryaznov, S.M. and Schultz, R.G., 2011. "2'-Arabino-Fluorooligonucleotide N3'-P5' Phosphoramidates: Their synthesis and use", US 2011/0294213 A1 (US) C12N 5/09, C07H 19/20, G01N 33/53, C070 21/02, C070 21/04, C070 19/04, C070 19/10

Gu, W., Zhou, T., and Wilke, C.O., 2010. "A universal trend of reduced mRNA stability near the translation-initiation site in prokaryotes and eukaryotes". PLoS Comput Biol, 6(2), e1000664

Guan, R., Ho, M.C., Almo, S.C., and Schramm, V.L., 2011. "Methylthioinosine Phosphorylase from Pseudomonas aeruginosa. Structure and Annotation of a Novel Enzyme in Quorum Sensing". Biochemistry-Us, 50(7), 1247-1254

Gura, T., 1995. "Antisense Has Growing Pains". Science, 270(5236), 575-577 Hagmann, W.K., 2008. "The many roles for fluorine in medicinal chemistry". J Med Chem, 51(15),

4359-4369 Hajdo, L., Szulc, A.B., Klajnert, B., and Bryszewska, M., 2010. "Metabolic Limitations of the Use of

Nucleoside Analogs in Cancer Therapy May Be Overcome by Application of Nanoparticles as Drug Carriers: A Review". Drug Develop Res, 71(7), 383-394

Haki, G.D. and Rakshit, S.K., 2003. "Developments in industrially important thermostable enzymes: a review". Bioresource Technol, 89(1), 17-34

Hamamoto, T., Noguchi, T., and Midorikawa, Y., 1996. "Purification and characterization of purine nucleoside phosphorylase and pyrimidine nucleoside phosphorylase from Bacillus stearothermophilus TH 6-2". Biosci Biotech Bioch, 60(7), 1179-1180

Hamamoto, T., Noguchi, T., and Midorikawa, Y., 1997a. "Cloning of purine nucleoside phosphorylase II gene from Bacillus stearothermophilus TH 6-2 and characterization of its gene product". Biosci Biotech Bioch, 61(2), 276-280

Hamamoto, T., Okuyama, K., Noguchi, T., and Midorikawa, Y., 1997b. "Cloning and expression of purine nucleoside phosphorylase I gene from Bacillus stearothermophilus TH 6-2". Biosci Biotech Bioch, 61(2), 272-275

Hammer-Jespersen, K. and Munch-Ptersen, A., 1975. "Multiple regulation of nucleoside catabolizing enzymes: regulation of the deo operon by the cytR and deoR gene products". Mol Gen Genet, 137(4), 327-335

Hammond, D.J. and Gutteridge, W.E., 1984. "Purine and Pyrimidine Metabolism in the Trypanosomatidae". Mol Biochem Parasit, 13(3), 243-261

Hammond, S.M., Bernstein, E., Beach, D., and Hannon, G.J., 2000. "An RNA-directed nuclease mediates post-transcriptional gene silencing in Drosophila cells". Nature, 404(6775), 293-296

Hatahet, F., Nguyen, V.D., Salo, K.E., and Ruddock, L.W., 2010. "Disruption of reducing pathways is not essential for efficient disulfide bond formation in the cytoplasm of E. coli". Microb Cell Fact, 967

Hempel, A., Camerman, N., Grierson, J., Mastropaolo, D., and Camerman, A., 1999. "FF-β-D-Arabinofuranosyluracil". Acta Crystallogr C, 55632-633

Hénaut, A. and Danchin, A., "Analysis and Predictions from Escherichia coli sequences", in "Escherichia coli and Salmonella typhimurium cellular and molecular biology", F. Neidhardt, et al., Editors. 1996, ASM press, Washington, DC. p. 2047 -2066.

Henne, A., Bruggemann, H., Raasch, C., Wiezer, A., Hartsch, T., Liesegang, H., Johann, A., Lienard, T., Gohl, O., Martinez-Arias, R., Jacobi, C., Starkuviene, V., Schlenczeck, S., Dencker, S., Huber, R., Klenk, H.P., Kramer, W., Merkl, R., Gottschalk, G., and Fritz, H.J., 2004. "The genome sequence of the extreme thermophile Thermus thermophilus". Nat Biotechnol, 22(5), 547-553

124 References

Hennen, W.J. and Wong, C.H., 1989. "A New Method for the Enzymatic-Synthesis of Nucleosides Using Purine Nucleoside Phosphorylase". J Org Chem, 54(19), 4692-4695

Hidalgo, A., Betancor, L., Moreno, R., Zafra, O., Cava, F., Fernandez-Lafuente, R., Guisan, J.M., and Berenguer, J., 2004. "Thermus thermophilus as a cell factory for the production of a thermophilic Mn-dependent catalase which fails to be synthesized in an active form in Escherichia coli". Appl Environ Microbiol, 70(7), 3839-3844

Hocek, M. and Dvorakova, H., 2003. "An efficient synthesis of 2-substituted 6-methylpurine bases and nucleosides by Fe- or Pd-catalyzed cross-coupling reactions of 2,6-dichloropurines". J Org Chem, 68(14), 5773-5776

Holguin, J. and Cardinaud, R., 1975. "Trans-N-Deoxyribosylase - Purification by Affinity Chromatography and Characterization". Eur J Biochem, 54(2), 505-514

Hori, N., Watanabe, M., Yamazaki, Y., and Mikami, Y., 1989a. "Purification and Characterization of 2nd Thermostable Purine Nucleoside Phosphorylase in Bacillus-Stearothermophilus Jts-859". Agr Biol Chem Tokyo, 53(12), 3219-3224

Hori, N., Watanabe, M., Yamazaki, Y., and Mikami, Y., 1989b. "Purification and Characterization of Thermostable Purine Nucleoside Phosphorylase of Bacillus-Stearothermophilus Jts-859". Agr Biol Chem Tokyo, 53(8), 2205-2210

Hori, N., Watanabe, M., Yamazaki, Y., and Mikami, Y., 1990. "Purification and Characterization of Thermostable Pyrimidine Nucleoside Phosphorylase from Bacillus-Stearothermophilus Jts-859". Agr Biol Chem Tokyo, 54(3), 763-768

Hori, N., Watanabe, M., Sunagawa, K., Uehara, K., and Mikami, Y., 1991. "Production of 5-methyluridine by immobilized thermostable purine nucleoside phosphorylase and pyrimidine nucleoside phosphorylase from Bacillus stearothermophilus JTS 859". J Biotechnol, 17(2), 121-131

Howell, H.G., Brodfuehrer, P.R., Brundidge, S.P., Benigni, D.A., and Sapino, C., 1988. "Antiviral Nucleosides - a Stereospecific, Total Synthesis of 2'-Fluoro-2'-Deoxy-β-D-Arabinofuranosyl Nucleosides". J Org Chem, 53(1), 85-88

Huang, M.C., Montgomery, J.A., Thorpe, M.C., Stewart, E.L., Secrist, J.A., and Blakley, R.L., 1983. "Formation of 3-(2'-Deoxyribofuranosyl) and 9-(2'-Deoxyribofuranosyl) Nucleosides of 8-Substituted Purines by Nucleoside Deoxyribosyltransferase". Arch Biochem Biophys, 222(1), 133-144

Ikemura, T., 1981. "Correlation between the Abundance of Escherichia coli Transfer-RNAs and the Occurrence of the Respective Codons in Its Protein Genes - a Proposal for a Synonymous Codon Choice That Is Optimal for the Escherichia coli Translational System". J Mol Biol, 151(3), 389-409

Ishii, M., Shirae, H., and Yokozeki, K., 1989. "Enzymatic Production of 5-Methyluridine from Purine Nucleosides and Thymine by Erwinia carotovora AJ 2992". Agr Biol Chem Tokyo, 53(12), 3209-3218

Ivanov, I., Alexandrova, R., Dragulev, B., Saraffova, A., and AbouHaidar, M.G., 1992. "Effect of tandemly repeated AGG triplets on the translation of CAT-mRNA in E. coli". FEBS Lett, 307(2), 173-176

Jensen, K.F. and Nygaard, P., 1975. "Purine Nucleoside Phosphorylase from Escherichia coli and Salmonella typhimurium - Purification and Some Properties". Eur J Biochem, 51(1), 253-265

Jensen, K.F., 1978. "Two purine nucleoside posphorylases in Bacillus subtilis. Purification and some properties of adenosine-specific phosphorylase". BBA-Enzymol, 525(2), 346-356

Jin, H., Zhao, Q., Gonzalez de Valdivia, E.I., Ardell, D.H., Stenstrom, M., and Isaksson, L.A., 2006. "Influences on gene expression in vivo by a Shine-Dalgarno sequence". Mol Microbiol, 60(2), 480-492

Jung, S.T., Kang, T.H., and Georgiou, G., 2010. "Efficient expression and purification of human aglycosylated Fcgamma receptors in Escherichia coli". Biotechnol Bioeng, 107(1), 21-30

Kalckar, H.M., 1947. "Differential spectrophotometry of purine compounds by means of specific enzymes; studies of the enzymes of purine metabolism". J Biol Chem, 167(2), 461-475

References 125

Kalota, A., Karabon, L., Swider, C.R., Viazovkina, E., Elzagheid, M., Damha, M.J., and Gewirtz, A.M., 2006. "2'-Deoxy-2'-fluoro-β-D-arabinonucleic acid (2'F-ANA) modified oligonucleotides (ON) effect highly efficient, and persistent, gene silencing". Nucleic Acids Res, 34(2), 451-461

Kaminski, P.A., 2002. "Functional cloning, heterologous expression, and purification of two different N-deoxyribosyltransferases from Lactobacillus helveticus". J Biol Chem, 277(17), 14400-14407

Kane, J.F., 1995. "Effects of Rare Codon Clusters on High-Level Expression of Heterologous Proteins in Escherichia coli". Curr Opin Biotech, 6(5), 494-500

Kaushik, J.K., Ogasahara, K., and Yutani, K., 2002. "The Unusually Slow Relaxation Kinetics of the Folding-unfolding of Pyrrolidone Carboxyl Peptidase from a Hyperthermophile, Pyrococcus furiosus". J Mol Biol, 316(4), 991-1003

Kawarabayasi, Y., Hino, Y., Horikawa, H., Yamazaki, S., Haikawa, Y., Jin-no, K., Takahashi, M., Sekine, M., Baba, S., Ankai, A., Kosugi, H., Hosoyama, A., Fukui, S., Nagai, Y., Nishijima, K., Nakazawa, H., Takamiya, M., Masuda, S., Funahashi, T., Tanaka, T., Kudoh, Y., Yamazaki, J., Kushida, N., Oguchi, A., Kikuchi, H., and et al., 1999. "Complete genome sequence of an aerobic hyper-thermophilic crenarchaeon, Aeropyrum pernix K1". DNA Res, 6(2), 83-101, 145-52

Kawasaki, A.M., Casper, M.D., Freier, S.M., Lesnik, E.A., Zounes, M.C., Cummins, L.L., Gonzalez, C., and Cook, P.D., 1993. "Uniformly Modified 2'-Deoxy-2'-Fluoro Phosphorothioate Oligonucleotides as Nuclease-Resistant Antisense Compounds with High-Affinity and Specificity for RNA Targets". J Med Chem, 36(7), 831-841

Kazimierczuk, Z., Cottam, H.B., Revankar, G.R., and Robins, R.K., 1984. "Synthesis of 2'-Deoxytubercidin, 2'-Deoxyadenosine, and Related 2'-Deoxynucleosides Via a Novel Direct Stereospecific Sodium-Salt Glycosylation Procedure". J Am Chem Soc, 106(21), 6379-6382

Khan, M.A., Sadaf, S., and Akhtar, M.W., 2007. "Role of Silent Gene Mutations in the Expression of Caprine Growth Hormone in Escherchia coli". Biotechnol Progr, 23(5), 1049-1052

Khati, M., Schuman, M., Ibrahim, J., Sattentau, Q., Gordon, S., and James, W., 2003. "Neutralization of infectivity of diverse R5 clinical isolates of human immunodeficiency virus type 1 by gp120-binding 2'F-RNA aptamers". J Virol, 77(23), 12692-12698

Koch, A.L. and Vallee, G., 1958. "The Properties of Adenosine Deaminase and Adenosine Nucleoside Phosphorylase in Extracts of Escherichia coli". J Biol Chem, 234(5), 1213-1218

Koma, D., Sawai, T., Harayama, S., and Kino, K., 2006. "Overexpression of the genes from thermophiles in Escherichia coli by high-temperature cultivation". Appl Microbiol Biotechnol, 73(1), 172-180

Koszalka, G.W., Vanhooke, J., Short, S.A., and Hall, W.W., 1988. "Purification and Properties of Inosine-Guanosine Phosphorylase from Escherichia Coli K12". J Bacteriol, 170(8), 3493-3498

Krahe, M., Antranikian, G., and Märkl, H., 1996. "Fermentation of extremophilic microorganisms". FEMS Microbiol Rev, 18271-285

Krause, M., Ukkonen, K., Haataja, T., Ruottinen, M., Glumoff, T., Neubauer, A., Neubauer, P., and Vasala, A., 2010. "A novel fed-batch based cultivation method provides high cell-density and improves yield of soluble recombinant proteins in shaken cultures". Microb Cell Fact, 9

Krenitsky, T.A., Koszalka, G.W., and Tuttle, J.V., 1981. "Purine nucleoside synthesis, an efficient method employing nucleoside phosphorylases". Biochemistry-Us, 20(12), 3615-3621

Krenitsky, T.A. and Tuttle, J.V., 1982. "Correlation of Substrate-Stabilization Patterns with Proposed Mechanisms for 3 Nucleoside Phosphorylases". BBA-Protein Struct M, 703(2), 247-249

Krzeminski, J., Nawrot, B., Pankiewicz, K.W., and Watanabe, K.A., 1991. "Nucleosides .157. Synthesis of 9-(2-Deoxy-2-Fluoro-β-D-Arabinofuranosyl)Hypoxanthine - the first Direct Introduction of a 2'-β-Fluoro Substituent in Preformed Purine Nucleosides - Studies Directed toward the Synthesis of 2'-Deoxy-2'-Substituted Arabinonucleosides ". Nucleos Nucleot, 10(4), 781-798

Kudla, G., Murray, A.W., Tollervey, D., and Plotkin, J.B., 2009. "Coding-sequence determinants of gene expression in Escherichia coli". Science, 324(5924), 255-258

Kurland, C. and Gallant, J., 1996. "Errors of heterologous protein expression". Curr Opin Biotech, 7(5), 489-493

126 References

Larkin, M.A., Blackshields, G., Brown, N.P., Chenna, R., McGettigan, P.A., McWilliam, H., Valentin, F., Wallace, I.M., Wilm, A., Lopez, R., Thompson, J.D., Gibson, T.J., and Higgins, D.G., 2007. "Clustal W and Clustal X version 2.0". Bioinformatics, 23(21), 2947-2948

Larson, E.T., Mudeppa, D.G., Gillespie, J.R., Mueller, N., Napuli, A.J., Arif, J.A., Ross, J., Arakaki, T.L., Lauricella, A., Detitta, G., Luft, J., Zucker, F., Verlinde, C.L., Fan, E., Van Voorhis, W.C., Buckner, F.S., Rathod, P.K., Hol, W.G., and Merritt, E.A., 2010. "The crystal structure and activity of a putative trypanosomal nucleoside phosphorylase reveal it to be a homodimeric uridine phosphorylase". J Mol Biol, 396(5), 1244-59

Lewkowicz, E.S., Martínez, N., Rogert, M.C., Porro, S., and Iribarren, A.M., 2000. "An improved microbial synthesis of purine nucleosides". Biotechnol Lett, 221277-1280

Li, X., Jiang, X., Li, H., and Ren, D., 2008. "Purine nucleoside phosphorylase from Pseudoalteromonas sp. Bsi590: molecular cloning, gene expression and characterization of the recombinant protein". Extremophiles, 12(3), 325-333

Ling, F., Inoue, Y., and Kimura, A., 1990. "Purification and Characterization of a Novel Nucleoside Phosphorylase from a Klebsiella Sp and Its Use in the Enzymatic Production of Adenine-Arabinoside". Appl Environ Microb, 56(12), 3830-3834

Liu, P., Sharon, A., and Chu, C.K., 2008. "Fluorinated Nucleosides: Synthesis and Biological Implication". J Fluor Chem, 129(9), 743-766

Liu, X., Kantarjian, H., and Plunkett, W., 2012. "Sapacitabine for cancer". Expert Opin Investig Drugs, 21(4), 541-555

Lowenthal, R.M. and Eaton, K., 1996. "Toxicity of chemotherapy". Hematol Oncol Clin North Am, 10(4), 967-990

Mahmoudian, M., Rudd, B.A.M., Cox, B., Drake, C.S., Hall, R.M., Stead, P., Dawson, M.J., Chandler, M., Livermore, D.G., Turner, N.J., and Jenkins, G., 1998. "A versatile procedure for the generation of nucleoside 5'-carboxylic acids using nucleoside oxidase". Tetrahedron, 54(28), 8171-8182

Mahmoudian, M., 2000. "Biocatalytic production of chiral pharmaceutical intermediates". Biocatal Biotransfor, 18105-118

Manoharan, M., Akinc, A., Pandey, R.K., Qin, J., Hadwiger, P., John, M., Mills, K., Charisse, K., Maier, M.A., Nechev, L., Greene, E.M., Pallan, P.S., Rozners, E., Rajeev, K.G., and Egli, M., 2011. "Unique Gene-Silencing and Structural Properties of 2'-Fluoro-Modified siRNAs". Angew Chem Int Edit, 50(10), 2284-2288

Mao, C., Cook, W.J., Zhou, M., Koszalka, G.W., Krenitsky, T.A., and Ealick, S.E., 1997. "The crystal structure of Escherichia coli purine nucleoside phosphorylase: a comparison with the human enzyme reveals a conserved topology". Structure, 5(10), 1373-1383

Marck, C., Lesyng, B., and Saenger, W., 1982. "The Crystal-Structures of 2'-Deoxy-2'-Fluorocytidine and 2'-Deoxy-2'-Fluorouridine". J Mol Struct, 82(1-2), 77-94

Martinez, J., Patkaniowska, A., Urlaub, H., Luhrmann, R., and Tuschl, T., 2002. "Single-stranded antisense siRNAs guide target RNA cleavage in RNAi". Cell, 110(5), 563-574

Maruyama, T., Takamatsu, S., Kozai, S., Satoh, Y., and Izawa, K., 1999. "Synthesis of 9-(2-deoxy-2-fluoro-β-D-arabinofuranosyl)adenine bearing a selectively removable protecting group". Chem Pharm Bull, 47(7), 966-970

Mcelwain, M.C., Williams, M.V., and Pollack, J.D., 1988. "Acholeplasma-Laidlawii B-Pg9 Adenine-Specific Purine Nucleoside Phosphorylase That Accepts Ribose-1-Phosphate, Deoxyribose-1-Phosphate, and Xylose-1-Phosphate". J Bacteriol, 170(2), 564-567

McNicholas, S., Potterton, E., Wilson, K.S., and Noble, M.E., 2011. "Presenting your structures: the CCP4mg molecular-graphics software". Acta Crystallogr D, 67(Pt 4), 386-94

Mendieta, J., Martin-Santamaria, S., Priego, E.M., Balzarini, J., Camarasa, M.J., Perez-Perez, M.J., and Gago, F., 2004. "Role of histidine-85 in the catalytic mechanism of thymidine phosphorylase as assessed by targeted molecular dynamics simulations and quantum mechanical calculations". Biochemistry-Us, 43(2), 405-414

Mikhailopulo, I.A., 2007. "Biotechnology of Nucleic Acid Constituents - State of the Art and Perspectives". Curr Org Chem, 11(4), 317-335

References 127

Mikhailopulo, I.A. and Miroshnikov, A.I., 2011. "Biologically important nucleosides: modern trends in biotechnology and application". Mendeleev Commun, 21(2), 57-68

Mitsuya, H., Weinhold, K.J., Furman, P.A., Stclair, M.H., Lehrman, S.N., Gallo, R.C., Bolognesi, D., Barry, D.W., and Broder, S., 1985. "3'-Azido-3'-Deoxythymidine (Bw A509u) - an Antiviral Agent That Inhibits the Infectivity and Cytopathic Effect of Human Lymphotropic-T Virus Type-Iii Lymphadenopathy-Associated Virus Invitro". P Natl Acad Sci USA, 82(20), 7096-7100

Monia, B.P., Lesnik, E.A., Gonzalez, C., Lima, W.F., Mcgee, D., Guinosso, C.J., Kawasaki, A.M., Cook, P.D., and Freier, S.M., 1993. "Evaluation of 2'-Modified Oligonucleotides Containing 2'-Deoxy Gaps as Antisense Inhibitors of Gene-Expression". J Biol Chem, 268(19), 14514-14522

Montgomery, J.A., Shortnacy, A.T., Carson, D.A., and Secrist, J.A., 1986. "9-(2-Deoxy-2-Fluoro-FF-β-D-Arabinofuranosyluracil-D-Arabinofuranosyl)Guanine - a Metabolically Stable Cytotoxic Analog of 2'-Deoxyguanosine". J Med Chem, 29(11), 2389-2392

Montgomery, J.A., Shortnacy-Fowler, A.T., Clayton, S.D., Riordan, J.M., and Secrist, J.A., 3rd, 1992. "Synthesis and biologic activity of 2'-fluoro-2-halo derivatives of 9-β-D-arabinofuranosyladenine". J Med Chem, 35(2), 397-401

Montilla Arevalo, R., Deroncelé Thomas, V.M., López Gómez, C., Pascual Gilabert, M., Estévez Company, C., and Castells Boliart, J., 2011. "Thermostable biocatalyst combination for nucleoside synthesis", EP 2 338 985 A1 (Institut Univ. de Ciència i Tecnologia, Barcelona, ES) C12N 9/10, C12P 19/38

Moore, D. and Dowhan, D., "Purification and concentration of DNA from aqueous solutions", in "Current protocols in molecular biology", F.M. Ausubel, et al., Editors. 2002, John Wiley & Sons, Houston. p. Unit 2.1A.

Moris, F. and Gotor, V., 1993. "A Useful and Versatile Procedure for the Acylation of Nucleosides through an Enzymatic-Reaction". J Org Chem, 58(3), 653-660

Morrissey, D.V., Lockridge, J.A., Shaw, L., Blanchard, K., Jensen, K., Breen, W., Hartsough, K., Machemer, L., Radka, S., Jadhav, V., Vaish, N., Zinnen, S., Vargeese, C., Bowman, K., Shaffer, C.S., Jeffs, L.B., Judge, A., MacLachlan, I., and Polisky, B., 2005. "Potent and persistent in vivo anti-HBV activity of chemically modified siRNAs". Nat Biotechnol, 23(8), 1002-1007

Müller, K., Faeh, C., and Diederich, F., 2007. "Fluorine in pharmaceuticals: looking beyond intuition". Science, 317(5846), 1881-1886

Murphy, E.L., Collier, A.C., Kalish, L.A., Assmann, S.F., Para, M.F., Flanigan, T.P., Kumar, P.N., Mintz, L., Wallach, F.R., Nemo, G.J., and Study, V.A.T., 2001. "Highly active antiretroviral therapy decreases mortality and morbidity in patients with advanced HIV disease". Ann Intern Med, 135(1), 17-26

Na, D., Lee, S., and Lee, D., 2010. "Mathematical modeling of translation initiation for the estimation of its efficiency to computationally design mRNA sequences with desired expression levels in prokaryotes". BMC Systems Biology, 4(1), 71

Nair, V., Adah, S.A., and Ha, S.B., 1995. "Strategically Functionalized Adenosines: Agonists for Adenosine Receptors". Nucleos Nucleot Nucl, 14(3-5), 537-553

Narberhaus, F., Waldminghaus, T., and Chowdhury, S., 2006. "RNA thermometers". FEMS Microbiology Reviews, 30(1), 3-16

Nazina, T.N., Tourova, T.P., Poltaraus, A.B., Novikova, E.V., Grigoryan, A.A., Ivanova, A.E., Lysenko, A.M., Petrunyaka, V.V., Osipov, G.A., Belyaev, S.S., and Ivanov, M.V., 2001. "Taxonomic study of aerobic thermophilic bacilli: descriptions of Geobacillus subterraneus gen. nov., sp nov and Geobacillus uzenensis sp nov from petroleum reservoirs and transfer of Bacillus stearothermophilus Bacillus thermocatenulatus, Bacillus thermoleovorans, Bacillus kaustophilus, Bacillus thermoglucosidasius and Bacillus thermodenitrificans to Geobacillus as the new combinations G. stearothermophilus, G. thermocatenulatus, G. thermoleovorans, G. kaustophilus, G. thermoglucosidasius and G. thermodenitrificans". Int J Syst Evol Micr, 51(Pt 2), 433-446

Neupert, J., Karcher, D., and Bock, R., 2008. "Design of simple synthetic RNA thermometers for temperature-controlled gene expression in Escherichia coli". Nucleic Acids Res, 36(19), e124

128 References

Ng, E.W.M. and Adamis, A.P., 2006. "Anti-VEGF aptamer (pegaptanib) therapy for ocular vascular diseases". Ann Ny Acad Sci, 1082151-171

Nguyen, V.D., Hatahet, F., Salo, K.E., Enlund, E., Zhang, C., and Ruddock, L.W., 2011. "Pre-expression of a sulfhydryl oxidase significantly increases the yields of eukaryotic disulfide bond containing proteins expressed in the cytoplasm of E.coli". Microb Cell Fact, 10(1), 1

Niemitalo, O., Neubauer, A., Liebal, U., Myllyharju, J., Juffer, A.H., and Neubauer, P., 2005. "Modelling of translation of human protein disulfide isomerase in Escherichia coli - A case study of gene optimisation". J Biotechnol, 120(1), 11-24

Nishihara, K., Kanemori, M., Kitagawa, M., Yanagi, H., and Yura, T., 1998. "Chaperone coexpression plasmids: Differential and synergistic roles of DnaK-DnaJ-GrpE and GroEL-GroES in assisting folding of an allergen of Japanese cedar pollen, Cryj2 in Escherichia coli". Appl Environ Microb, 64(5), 1694-1699

Oganesyan, N., Ankoudinova, I., Kim, S.H., and Kim, R., 2007. "Effect of osmotic stress and heat shock in recombinant protein overexpression and crystallization". Protein Expr Purif, 52(2), 280-285

Ogasahara, K., Nakamura, M., Nakura, S., Tsunasawa, S., Kato, I., Yoshimoto, T., and Yutani, K., 1998. "The Unusually Slow Unfolding Rate Causes the High Stability of Pyrrolidone Carboxyl Peptidase from a Hyperthermophile, Pyrococcus furiosus: Equilibrium and Kinetic Studies of Guanidine Hydrochloride-Induced Unfolding and Refolding". Biochemistry-Us, 37(50), 17537-17544

Ohrui, H., 2011. "Development of modified nucleosides that have supremely high anti-HIV activity and low toxicity and prevent the emergence of resistant HIV mutants". P Jpn Acad B-Phys, 87(3), 53-65

Okuyama, K., Shibuya, S., Hamamoto, T., and Noguchi, T., 2003. "Enzymatic synthesis of 2'-deoxyguanosine with nucleoside deoxyribosyltransferase-II". Biosci Biotech Bioch, 67(5), 989-995

Oshima, T. and Imahori, K., 1974. "Description of Thermus Thermophilus (Yoshida and Oshima) comb. nov, a Nonsporulating Thermophilic Bacterium from a Japanese Thermal Spa". Int J Syst Bacteriol, 24(1), 102-112

Pace, C.N., Vajdos, F., Fee, L., Grimsley, G., and Gray, T., 1995. "How to Measure and Predict the Molar Absorption-Coefficient of a Protein". Protein Sci, 4(11), 2411-2423

Panova, N.G., Shcheveleva, E.V., Alexeev, C.S., Mukhortov, V.G., Zuev, A.N., Mikhailov, S.N., Esipov, R.S., Chuvikovsky, D.V., and Miroshnikov, A.I., 2004. "Use of 4-thiouridine and 4-thiothymidine in studies on pyrimidine nucleoside phosphorylases". Mol Biol+, 38(5), 770-776

Panova, N.G., Alexeev, C.S., Kuzmichov, A.S., Shcheveleva, E.V., Gavryushov, S.A., Polyakov, K.M., Kritzyn, A.M., Mikhailov, S.N., Esipov, R.S., and Miroshnikov, A.I., 2007. "Substrate specificity of Escherichia coli thymidine phosphorylase". Biochemistry (Mosc), 72(1), 21-28

Panula-Perälä, J., Šiurkus, J., Vasala, A., Wilmanowski, R., Casteleijn, M.G., and Neubauer, P., 2008. "Enzyme controlled glucose auto-delivery for high cell density cultivations in microplates and shake flasks". Microb Cell Fact, 7(31),

Parker, W.B., King, S.A., Allan, P.W., Bennett, L.L., Secrist, J.A., Montgomery, J.A., Gilberg, K.S., Waud, W.R., Wells, A.H., Gillespie, G.Y., and Sorscher, E.J., 1997. "In vivo gene therapy of cancer with E. coli purine nucleoside phosphorylase". Hum Gene Ther, 8(14), 1637-1644

Parker, W.B., Allan, P.W., Hassan, A.E., Secrist, J.A., 3rd, Sorscher, E.J., and Waud, W.R., 2003. "Antitumor activity of 2-fluoro-2'-deoxyadenosine against tumors that express Escherichia coli purine nucleoside phosphorylase". Cancer Gene Ther, 10(1), 23-9

Poulsen, S.A. and Quinn, R.J., 1998. "Adenosine receptors: new opportunities for future drugs". Bioorg Med Chem, 6(6), 619-641

Pugmire, M.J. and Ealick, S.E., 1998. "The crystal structure of pyrimidine nucleoside phosphorylase in a closed conformation". Structure, 6(11), 1467-1479

Pugmire, M.J. and Ealick, S.E., 2002. "Structural analyses reveal two distinct families of nucleoside phosphorylases". Biochemical Journal, 3611-25

References 129

Ranganathan, R., 1977. "Modification of 2'-Position of Purine Nucleosides - Syntheses of 2'-α-Substituted-2'-Deoxyadenosine Analogs". Tetrahedron Lett, (15), 1291-1294

Rentmeister, A., Arnold, F.H., and Fasan, R., 2009. "Chemo-enzymatic fluorination of unactivated organic compounds". Nat Chem Biol, 5(1), 26-28

Rocchietti, S., Ubiali, D., Terreni, M., Albertini, A.M., Fernandez-Lafuente, R., Guisan, J.M., and Pregnolato, M., 2004. "Immobilization and stabilization of recombinant multimeric uridine and purine nucleoside phosphorylases from Bacillus subtilis". Biomacromolecules, 5(6), 2195-2200

Rodenko, B., van der Burg, A.M., Wanner, M.J., Kaiser, M., Brun, R., Gould, M., de Koning, H.P., and Koomen, G.J., 2007. "2,N6-disubstituted adenosine analogs with antitrypanosomal and antimalarial activities". Antimicrob Agents Chemother, 51(11), 3796-3802

Rogert, M.C., Trelles, J.A., Porro, S., Lewkowicz, E.S., and Iribarren, A.M., 2002. "Microbial synthesis of antiviral nucleosides using Escherichia coli BL21 as biocatalyst". Biocatal Biotransfor, 20(5), 347-351

Roosild, T.P., Castronovo, S., Fabbiani, M., and Pizzorno, G., 2009. "Implications of the structure of human uridine phosphorylase 1 on the development of novel inhibitors for improving the therapeutic window of fluoropyrimidine chemotherapy". Bmc Struct Biol, 9

Roshevskaia, L.A., Barai, V.N., Zinchenko, A.I., Kvasiuk, E.I., and Mikhailopulo, I.A., 1986. "[Preparative synthesis of the antiviral nucleoside 9-β-D-arabinofuranosyladenine by using bacterial cells]". Antibiot Med Biotekhnol, 31(3), 174-178

Sadaf, S., Khan, M.A., and Akhtar, M.W., 2008. "Expression enhancement of bubaline somatotropin in E. coli through gene modifications in the 5'-end coding region". J Biotechnol, 135(2), 134-139

Sako, Y., Nomura, N., Uchida, A., Ishida, Y., Morii, H., Koga, Y., Hoaki, T., and Maruyama, T., 1996. "Aeropyrum pernix gen. nov., sp. nov., a novel aerobic hyperthermophilic archaeon growing at temperatures up to 100 °C". Int J Syst Bacteriol, 46(4), 1070-1077

Salvatori, D., Volpini, R., Vincenzetti, S., Vita, A., Costanzi, S., Lambertucci, C., Cristalli, G., and Vittori, S., 2002. "Adenine and deazaadenine nucleoside and deoxynucleoside analogues: inhibition of viral replication of sheep MVV (in vitro model for HIV) and bovine BHV-1". Bioorg Med Chem, 10(9), 2973-80

Sambrook, J. and Russell, D.W., "Molecular cloning : a laboratory manual". 3rd ed. 2001, New York: Cold Spring Harbor Laboratory Press.

Sampath, D., Rao, V.A., and Plunkett, W., 2003. "Mechanisms of apoptosis induction by nucleoside analogs". Oncogene, 22(56), 9063-9074

Samsel, M. and Dzierzbicka, K., 2011. "Therapeutic potential of adenosine analogues and conjugates". Pharmacol Rep, 63(3), 601-617

Santaniello, E., Ciuffreda, P., and Alessandrini, L., 2005. "Synthesis of modified purine nucleosides and related compounds mediated by adenosine deaminase (ADA) and adenylate deaminase (AMPDA)". Synthesis-Stuttgart, (4), 509-526

Saunders, P.P., Wilson, B.A., and Saunders, G.F., 1969. "Purification and comparative properties of a pyrimidine nucleoside phosphorylase from Bacillus stearothermophilus". J Biol Chem, 244(13), 3691-3697

Schaeffer, H.J. and Thomas, H.J., 1958. "Synthesis of Potential Anticancer Agents. XIV. Ribosides of 2,6-Disubstituted Purines". J Am Chem Soc, 80(14), 3738-3742

Scheer, M., Grote, A., Chang, A., Schomburg, I., Munaretto, C., Rother, M., Sohngen, C., Stelzer, M., Thiele, J., and Schomburg, D., 2011. "BRENDA, the enzyme information system in 2011". Nucleic Acids Res, 39D670-D676

Schein, C.H. and Noteborn, M.H.M., 1988. "Formation of Soluble Recombinant Proteins in Escherichia coli Is Favored by Lower Growth Temperature". Bio-Technol, 6(3), 291-294

Schlieker, C., Bukau, B., and Mogk, A., 2002. "Prevention and reversion of protein aggregation by molecular chaperones in the E. coli cytosol: implications for their applicability in biotechnology". J Biotechnol, 96(1), 13-21

Schnick, C., Robien, M.A., Brzozowski, A.M., Dodson, E.J., Murshudov, G.N., Anderson, L., Luft, J.R., Mehlin, C., Hol, W.G.J., Brannigan, J.A., and Wilkinson, A.J., 2005. "Structures of Plasmodium

130 References

falciparum purine nucleoside phosphorylase complexed with sulfate and its natural substrate inosine". Acta Crystallogr D, 61(Pt 9), 1245-1254

Schultheisz, H.L., Szymczyna, B.R., Scott, L.G., and Williamson, J.R., 2008. "Pathway engineered enzymatic de novo purine nucleotide synthesis". ACS Chem Biol, 3(8), 499-511

Schultheisz, H.L., Szymczyna, B.R., Scott, L.G., and Williamson, J.R., 2010. "Enzymatic De Novo Pyrimidine Nucleotide Synthesis". J Am Chem Soc, 133(2), 297-304

Serova, M., Galmarini, C.M., Ghoul, A., Benhadji, K., Green, S.R., Chiao, J., Faivre, S., Cvitkovic, E., Le Tourneau, C., Calvo, F., and Raymond, E., 2007. "Antiproliferative effects of sapacitabine (CYC682), a novel 2'-deoxycytidine-derivative, in human cancer cells". Brit J Cancer, 97(5), 628-636

Serra, I., Serra, C.D., Rocchietti, S., Ubiali, D., and Terreni, M., 2011. "Stabilization of thymidine phosphorylase from Escherichia coli by immobilization and post immobilization techniques". Enzyme Microb Tech, 49(1), 52-58

Sgarrella, F., Poddie, F.P.A., Meloni, M.A., Sciola, L., Pippia, P., and Tozzi, M.G., 1997. "Channelling of deoxyribose moiety of exogenous DNA into carbohydrate metabolism: Role of deoxyriboaldolase". Comp Biochem Phys B, 117(2), 253-257

Sgarrella, F., Frassetto, L., Allegnini, S., Camici, M., Carta, M.C., Fadda, P., Tozzi, M.G., and Ipata, P.L., 2007. "Characterization of the adenine nucleoside specific phosphorylase of Bacillus cereus". Bba-Gen Subjects, 1770(10), 1498-1505

Sharp, P.M. and Li, W.H., 1987. "The codon Adaptation Index -a measure of directional synonymous codon usage bias, and its potential applications". Nucleic Acids Res, 15(3), 1281-1295

Shi, J.X., Du, J.F., Ma, T.W., Pankiewicz, K.W., Patterson, S.E., Tharnish, P.M., McBrayer, T.R., Stuyver, L.J., Otto, M.J., Chu, C.K., Schinazi, R.F., and Watanabe, K.A., 2005. "Synthesis and anti-viral activity of a series of D- and L-2'-deoxy-2'-fluororibonucleosides in the subgenomic HCV replicon system". Bioorgan Med Chem, 13(5), 1641-1652

Shi, W.X., Ting, L.M., Kicska, G.A., Lewandowicz, A., Tyler, P.C., Evans, G.B., Furneaux, R.H., Kim, K., Almo, S.C., and Schramm, V.L., 2004. "Plasmodium falciparum purine nucleoside phosphorylase - Crystal structures, immucillin inhibitors, and dual catalytic function". J Biol Chem, 279(18), 18103-18106

Shiloach, J. and Fass, R., 2005. "Growing E. coli to high cell density - A historical perspective on method development". Biotechnol Adv, 23(5), 345-357

Shimizu, K. and Kunishima, N., 2007. "Purification, crystallization and preliminary X-ray diffraction study on pyrimidine nucleoside phosphorylase TTHA1771 from Thermus thermophilus HB8". Acta Crystallogr F, 63308-310

Shirae, H. and Yokozeki, K., 1991. "Purifications and Properties of Orotidine-Phosphorolyzing Enzyme and Purine Nucleoside Phosphorylase from Erwinia carotovora AJ 2992". Agric Biol Chem 55(7), 1849-1857

Silva, R.G., Nunes, J.E.S., Canduri, F., Borges, J.C., Gava, L.M., Moreno, F.B., Basso, L.A., and Santos, D.S., 2007. "Purine nucleoside phosphorylase: A potential target for the development of drugs to treat T-cell- and apicomplexan parasite-mediated diseases". Curr Drug Targets, 8(3), 413-422

Šiurkus, J., Panula-Perälä, J., Horn, U., Kraft, M., Rimseliene, R., and Neubauer, P., 2010. "Novel approach of high cell density recombinant bioprocess development: Optimisation and scale-up from microlitre to pilot scales while maintaining the fed-batch cultivation mode of E. coli cultures". Microb Cell Fact, 9(1), 35

Sivets, G.G., Kalinichenko, E.N., and Mikhailopulo, I.A., 2006. "Synthesis of C2'-ß-fluoro-substituted adenine nucleosides via pivaloyl derivatives of adenosine and 3'-deoxyadenosine". Lett Org Chem, 3(5), 402-408

Sørensen, H.P. and Mortensen, K.K., 2005. "Advanced genetic strategies for recombinant protein expression in Escherichia coli". J Biotechnol, 115(2), 113-128

Stavrovskaya, A.A., 2000. "Cellular mechanisms of multidrug resistance of tumor cells". Biochemistry (Mosc), 65(1), 95-106

References 131

Stephenson, M.L. and Zamecnik, P.C., 1978. "Inhibition of Rous sarcoma viral RNA translation by a specific oligodeoxyribonucleotide". P Natl Acad Sci USA, 75(1), 285-288

Suzuki, Y., Kishigami, T., Inoue, K., Mizoguchi, Y., Eto, N., Takagi, M., and Abe, S., 1983. "Bacillus thermoglucosidasius sp. nov., a New Species of Obligately Thermophilic Bacilli". Syst Appl Microbiol, 4(4), 487-495

Szeker, K., Niemitalo, O., Casteleijn, M.G., Juffer, A.H., and Neubauer, P., 2011. "High-temperature cultivation and 5'mRNA optimization are key factors for the efficient overexpression of thermostable Deinococcus geothermalis purine nucleoside phosphorylase in Escherichia coli". J Biotechnol, 156(4), 268-274

Szeker, K., Zhou, X., Schwab, T., Casanueva, A., Cowan, D., Mikhailopulo, I.A., and Neubauer, P., 2012. "Comparative investigations on thermostable pyrimidine nucleoside phosphorylases from Geobacillus thermoglucosidasius and Thermus thermophilus". J Mol Catal B-Enzym, in press(doi:10.1016/j.molcatb.2012.02.006),

Tahirov, T.H., Inagaki, E., Ohshima, N., Kitao, T., Kuroishi, C., Ukita, Y., Takio, K., Kobayashi, M., Kuramitsu, S., Yokoyama, S., and Miyano, M., 2004. "Crystal structure of purine nucleoside phosphorylase from Thermus thermophilus". J Mol Biol, 337(5), 1149-1160

Tann, C.H., Brodfuehrer, P.R., Brundidge, S.P., Sapino, C., and Howell, H.G., 1985. "Fluorocarbohydrates in Synthesis. An Efficient Synthesis of 1-(2-Deoxy-2-Fluoro-β-D-Arabinofuranosyl)-5-Iodouracil (β-FIAU) and 1-(2-Deoxy-2-Fluoro-β-D-Arabinofuranosyl)Thymine (β-FMAU)". J Org Chem, 50(19), 3644-3647

Taran, S.A., Verevkina, K.N., Feofanov, S.A., and Miroshnikov, A.I., 2009. "Enzymatic Transglycosylation of Natural and Modified Nucleosides by Immobilized Thermostable Nucleoside Phosphorylases from Geobacillus stearothermophilus". Russ J Bioorg Chem+, 35(6), 739-745

Tebbe, J., Wielgus-Kutrowska, B., Schroder, W., Luic, M., Shugar, D., Saenger, W., Koellner, G., and Bzowska, A., 1997. "Purine nucleoside phosphorylase (PNP) from Cellulomonas sp., a third class of PNP different from both ''low-molecular weight'' mammalian and ''high-molecular weight'' bacterial PNPs". Protein Eng, 1090

Tebbe, J., Bzowska, A., Wielgus-Kutrowska, B., Schroder, W., Kazimierczuk, Z., Shugar, D., Saenger, W., and Koellner, G., 1999. "Crystal structure of the purine nucleoside phosphorylase (PNP) from Cellulomonas sp and its implication for the mechanism of trimeric PNPs". J Mol Biol, 294(5), 1239-1255

Tennilä, T., Azhayeva, E., Vepsalainen, J., Laatikainen, R., Azhayev, A., and Mikhailopulo, I.A., 2000. "Oligonucleotides containing 9-(2-deoxy-2-fluoro-β-D-arabinofuranosyl)-adenine and -guanine: Synthesis, hybridization and antisense properties". Nucleos Nucleot Nucl, 19(10-12), 1861-1884

Thiel, A., Pries, R., Jeske, S., Trenkle, T., and Wollenberg, B., 2009. "Effect of Head and Neck Cancer Supernatant and CpG-Oligonucleotides on Migration and IFN-alpha Production of Plasmacytoid Dendritic Cells". Anticancer Res, 29(8), 3019-3025

Thomas, H.J., Tiwari, K.N., Clayton, S.J., Secrist, J.A., and Montgomery, J.A., 1994. "Synthesis and Biologic Activity of Purine 2’-Deoxy-2’-fluoro-ribonucleosides". Nucleos Nucleot, 3(1-3), 309-323

Tozzi, M.G., Sgarrella, F., and Ipata, P.L., 1981. "Induction and Repression of Enzymes Involved in Exogenous Purine Compound Utilization in Bacillus cereus". Bba-Gen Subjects, 678(3), 460-466

Trelles, J.A., Valino, A.L., Runza, V., Lewkowicz, E.S., and Iribarren, A.M., 2005. "Screening of catalytically active microorganisms for the synthesis of 6-modified purine nucleosides". Biotechnol Lett, 27(11), 759-763

Trembacz, H. and Jezewska, M.M., 1993. "Specific Adenosine Phosphorylase from Hepatopancreas of Gastropod Helix pomatia". Comp Biochem Phys B, 104(3), 481-487

Turkman, N., Gelovani, J.G., and Alauddin, M.M., 2010. "A novel method for stereospecific fluorination at the 2'-arabino-position of pyrimidine nucleoside: synthesis of [(18)F]-FMAU". J Labelled Compd Rad, 53(13-14), 782-786

132 References

Tuttle, J.V. and Krenitsky, T.A., 1992. "Therapeutic nucleosides ", EP 0 285 432 B1 (Wellcome Found, GB) C07H 19/16

Tuttle, J.V., Tisdale, M., and Krenitsky, T.A., 1993. "Purine 2'-deoxy-2'-fluororibosides as antiinfluenza virus agents". J Med Chem, 36(1), 119-125

Ubiali, D., Serra, C.D., Serra, I., Morelli, C.F., Terreni, M., Albertini, A.M., Manitto, P., and Speranza, G., 2012. "Production, Characterization and Synthetic Application of a Purine Nucleoside Phosphorylase from Aeromonas hydrophila". Adv Synth Catal, 354(1), 96-104

Ukkonen, K., Vasala, A., Ojamo, H., and Neubauer, P., 2011. "High-yield production of biologically active recombinant protein in shake flask culture by combination of enzyme-based glucose delivery and increased oxygen transfer". Microb Cell Fact, 10(107),

Utagawa, T., Morisawa, H., Miyoshi, T., Yoshinaga, F., Yamazaki, A., and Mitsugi, K., 1980. "A novel and simple method for the preparation of adenine arabinoside by bacterial transglycosylation reaction". FEBS Lett, 109(2), 261-263

Utagawa, T., Morisawa, H., Shigeru, Y., Yamazaki, A., Yoshinaga, F., and Hirose, Y., 1985a. "Properties of Nucleoside Phosphorylase from Enterobacter aerogenes". Agric Biol Chem, 49(11), 3239-3246

Utagawa, T., Morisawa, H., Yoshinaga, F., Yamazaki, A., Mitsugi, K., and Hirose, Y., 1985b. "Enzymatic-Synthesis of Nucleoside Antibiotics .1. Microbiological Synthesis of Adenine-Arabinoside". Agric Biol Chem, 49(4), 1053-1058

Utagawa, T., 1999. "Enzymatic preparation of nucleoside antibiotics". J Mol Catal B-Enzym, 6(3), 215-222

Visser, D., Hennessy, F., Rashamuse, K., Louw, M., and Brady, D., 2010. "Cloning, purification and characterisation of a recombinant purine nucleoside phosphorylase from Bacillus halodurans Alk36". Extremophiles, 14(2), 185-192

Visser, D.F., Hennessy, F., Rashamuse, J., B., P., and Brady, D., 2011. "Stabilization of Escherichia coli uridine phosphorylase by evolution and immobilization". J Mol Catal B-Enzym, 68279-285

Vita, A., Huang, C.Y., and Magni, G., 1983. "Uridine Phosphorylase from Escherichia coli B - Kinetic Studies on the Mechanism of Catalysis". Arch Biochem Biophys, 226(2), 687-692

Vittori, S., Lorenzen, A., Stannek, C., Costanzi, S., Volpini, R., AP, I.J., Kunzel, J.K., and Cristalli, G., 2000. "N-cycloalkyl derivatives of adenosine and 1-deazaadenosine as agonists and partial agonists of the A(1) adenosine receptor". J Med Chem, 43(2), 250-260

Vorbrüggen, H. and Ruh-Pohlenz, C., "Handbook of Nucleoside Synthesis". 1 ed. 2001: Wiley-Interscience.

Wang, Y. and Zhang, Y.H., 2009. "Overexpression and simple purification of the Thermotoga maritima 6-phosphogluconate dehydrogenase in Escherichia coli and its application for NADPH regeneration". Microb Cell Fact, 8(30),

Watanabe, K.A., Reichman, U., Hirota, K., Lopez, C., and Fox, J.J., 1979. "Nucleosides. 110. Synthesis and antiherpes virus activity of some 2'-fluoro-2'-deoxyarabinofuranosylpyrimidine nucleosides". J Med Chem, 22(1), 21-24

Watts, J.K. and Damha, M.J., 2008. "2'F-arabinonucleic acids (2'F-ANA) - History, properties, and new frontiers". Can J Chem, 86(7), 641-656

Watts, J.K. and Corey, D.R., 2010. "Clinical status of duplex RNA". Bioorg Med Chem Lett, 20(11), 3203-3207

Watts, J.K. and Corey, D.R., 2012. "Silencing disease genes in the laboratory and the clinic". J Pathol, 226(2), 365-379

Wei, L., Altman, R.B., and Chang, J.T., 1997. "Using the radial distributions of physical features to compare amino acid environments and align amino acid sequences". Pac Symp Biocomput, 465-476

Wei, X.K., Ding, Q.B., Zhang, L., Guo, Y.L., Ou, L., and Wang, C.L., 2008. "Induction of nucleoside phosphorylase in Enterobacter aerogenes and enzymatic synthesis of adenine arabinoside". J Zhejiang Univ Sci B, 9(7), 520-526

Welch, M., Villalobos, A., Gustafsson, C., and Minshull, J., 2009. "You're one in a googol: optimizing genes for protein expression". J R Soc Interface, 6

References 133

Wright, G.E., Hildebrand, C., Freese, S., Dudycz, L.W., and Kazimierczuk, Z., 1987. "Convenient Synthesis of 2-Halo-2'-Deoxyadenosines". J Org Chem, 52(20), 4617-4618

Wright, J.A., Taylor, N.F., and Fox, J.J., 1969. "Nucleosides .60. Fluorocarbohydrates .22. Synthesis of 2-Deoxy-2-Fluoro-D-Arabinose and 9-(2-Deoxy-2-Fluoro-α- and -β-D-Arabinofuranosyl)Adenines". J Org Chem, 34(9), 2632-2636

Xie, X.X., Xia, J.G., He, K.F., Lu, L.N., Xu, Q.Y., and Chen, N., 2011. "Low-molecular-mass purine nucleoside phosphorylase: characterization and application in enzymatic synthesis of nucleoside antiviral drugs". Biotechnol Lett, 33(6), 1107-1112

Yamada, K., Matsumoto, N., and Hayakawa, H., 2009. "Stereoselective Synthesis of 2-Deoxy-2-Fluroarabinofuranosyl-α-1-phosphate and its Application to the Synthesis of 2'-Deoxy-2'-Fluoroarabinofuranosyl Purine Nucleosides by a chemo-enzymatic method". Nucleos Nucleot Nucl, 28(11-12), 1117-1130

Yamazaki, S., Yamazaki, J., Nishijima, K., Otsuka, R., Mise, M., Ishikawa, H., Sasaki, K., Tago, S., and Isono, K., 2006. "Proteome analysis of an aerobic hyperthermophilic crenarchaeon, Aeropyrum pernix K1". Mol Cell Proteomics, 5(5), 811-823

Zaitseva, G.V., Zinchenko, A.I., Barai, V.N., Pavlova, N.I., Boreko, E.I., and Mikhailopulo, I.A., 1999. "Chemical and enzymatic synthesis and antiviral properties of 2'-deoxy-2'-fluoroguanosine". Nucleos Nucleot, 18(4-5), 687-688

Zamecnik, P.C. and Stephenson, M.L., 1978. "Inhibition of Rous sarcoma virus replication and cell transformation by a specific oligodeoxynucleotide". P Natl Acad Sci USA, 75(1), 280-284

Zeng, Q., Huang, B., Danielsen, K., Shukla, R., and Nagy, T., 2004. "Facile and Practical Synthesis of 2,6-Dichloropurine". Org Process Res Dev, 8(6), 962-963

Zhu, S.Z., Ren, L., Wang, J.J., Zheng, G.J., and Tang, P.W., 2012. "Two-step efficient synthesis of 5-methyluridine via two thermostable nucleoside phosphorylase from Aeropyrum pernix". Bioorg Med Chem Lett, 22(5), 2102-2104

Zinchenko, A.I., Barai, V.N., Bokut, S.B., Kvasyuk, E.I., and Mikhailopulo, I.A., 1990. "Synthesis of 9-(β-D-arabinofuranosyl)guanine using whole cells of Escherichia coli". Appl Microbiol Biotechnol, 32(6), 658-661

Zuffi, G., Ghisotti, D., Oliva, I., Capra, E., Frascotti, G., Tonon, G., and Orsini, G., 2004. "Immobilized biocatalysts for the production of nucleosides and nucleoside analogues by enzymatic transglycosylation reactions". Biocatal Biotransfor, 22(1), 25-33

Zuker, M., 2003. "Mfold web server for nucleic acid folding and hybridization prediction". Nucleic Acids Res, 31(13), 3406-3415

Appendix

DgPNP gene sequence (Dgeo1497, 786 bp)

gtggtggcgcgtgtaccggcaaggcctttcgcttccccgcctgctaccctggaccgcgtgagtgtccacctgaatgcccggcccggcgaaattgcc

gaaactgtcctgcttcccggcgaccccctgcgcgcgcagcacatcgcggagacgttcttcgagaacccggtgcagcataacagcgtgcgcggcat

gctgggctttaccggcacgtaccggggcgtccccgtcagcgtgcagggcactggtatgggaatcgcctcatccatgatctacgtcaacgagctgat

ccgagactacggttgccagaccctgattcgcgtcggaacggcaggcagctatcagccggacgtgcacgtgcgcgacctcgtcctggcacaggcc

gcctgcactgacagcaatatcaacaacatccgcttcggtctgcggaacttcgccccgattgcggactttgagctgctgctgcgcgcctaccaaatg

gcccgggaccgtggcttcgcgacccatgtcggcaacatcctgtcctcggacaccttctatcaggacgatccggagagctacaaactctgggcgca

gtacggcgtgctggcggtggagatggaagccgccgggctgtataccctcgccgccaagtacggcgtgcgcgcgctcaccatcctgaccatctccg

accacctcgtcacacgggaggaaaccaccgctgaggagcggcagacgacatttaacggcatgattgaggtcgcgctggacgccgcactgggac

tggcagtaccttccaacagcatgtga

ApMTAP gene sequence (APE0993.1, 735 bp)

ttgaggaagccggttcacctcgaggcagggcccggcgacgtggcaccactcgtggtggcagtcggcgacccggggagggctgagaggctggcc

acaggcctcctcgaggacgctaggctggtgtcctccgccaggggtctgaaggtatacacgggcagcttcaacggctcggaggtaacgatagcca

cccacggcataggaggcccctcggcggctgtagtcttcgaggagctgaggatgcttggggctgaggttctggtgaggctcggcacctcgggcggc

ctatccaaggacctcaggctgggggacgtggtggtcgccgcgggggcgggctgctactggggctccgggggtagcatccagtacgcgggcgag

aggcccatgtgcctcccggcctcccccgaccccatattgacggcggggatatacaggggcctctcctccaggctcggggatagggtggttctagc

ccccgtcatgtcgagcgacgccttctacgccgagacgcccgaggctgccgggaggtggaggagcctcggcatggcggctgttgagatggagctc

cacaccctcttcagcatatcctggatacggggcttccgctcggccggggtgctcatagtctccgacctcctcctacccgagggtttcaaacgcatca

caccaggggagcttgcaaggagggaggttgaggttggcagggctctcctcgaggtcctcacaggaggagtctag

GtPNP gene sequence (ADNQ01000003.1, 708 bp)

ttgagcatccatatcgaagcaaagcaacaagaaattgctgagaaaattttgctgccgggtgatccgctccgcgcccaatatatcgcggaaacatt

tttagaaggagcgacatgctacaatcgcgtgcgcggcatgcttggatttacaggtacatacaaagggcaccgcatttctgtgcaagggacagga

atgggagttccttcgatttcgatttatgtcaatgaattaatccaaagctatcatgtacaaacactcattcgcgttggaacgtgcggagcgattcaaa

aagatgtcaacgttcgcgacgttattttggcgatgagcgcatccacggattcgaacatgaaccgtttgactttccgtgggcgggattatgcgccga

cagcaaattttgctttgttgcggaccgcttatgaagtcggcgccgaaaaagggcttccgctaaaggtcggaagtgtctttactgctgacatgtttta

taatgacgaaccggactgggagacgtgggcccgctacggtgttttagctgttgagatggaaacagcggcgctgtatacgctggcggcaaaatttg

gccgaaaagcgctttctgtgctgacggtaagcgaccatattttaacaggagaagaaacaacggcacaagaacggcaaacaacgtttaacgata

tgattgaagtggcgctggaaacggcgattcgcgtagaataa

GtPyNP gene sequence (NZ_ADNQ01000001.1, 1296 bp)

atggtcgatttaattgcgaaaaaacgggatggttatgagctttcaaaagaagaaattgattttatcatccgcggttacacgaacggcgacattcct

gattatcaaatgagcgcgttcgcgatggcggtgtttttccgcggcatgacagaagaagaaacggcggcgctaacgatggcgatggtccgctccg

gagatgtcatcgatttatcgaaaatcgaaggaatgaaagtcgacaagcattcaacgggtggcgtcggcgatacgacgacgcttgtgttagggcc

gcttgttgcgtctgtcggcgtgcctgtcgcaaaaatgtcggggcgagggcttggacataccggcgggaccattgataaattagagtccgttccagg

gtttcatgtggaaatcgataacgagcaatttattgagcttgtgaataaaaacaaaatcgcgattatcggccagacaggcaatttaacgccagccg

ataaaaagctgtatgcgctccgtgacgttacggcgacggtggacagcattccactgatcgcttcgtcgattatgagcaaaaaaattgccgctggc

136 Appendix

gctgacgcgattgtgttggatgtgaaaacgggagccggcgcgtttatgaaagattttgcaggagcgaagcggctcgcaacagcgatggtggaaa

tcggcaagcgcgtcggccggaaaacgatggcggttatttccgacatgagccagccgctcggatacgctgttggaaacgcgctcgaagtgaaaga

agcgattgatacgcttaaaggaaaagggccagaagatttacaagaactatgtttgacgcttggaagctatatggtatatttggcggaaaaagcct

cttcattagaggaagcgcgcgcgctgttagaagcgtcgattcgggaaggaaaagcgttagaaacgttcaaagtgtttctcagcgcgcaaggcgg

cgacgcatcggttgtcgatgatccaacgaaactgccgcaagcgaaataccgatgggagcttgaagccccggaagatgggtatgtcgcggaaatt

gtcgctgacgaagtcggaacggctgcgatgctgcttggagccgggcgggcgacaaaagaagcaacgatcgatctttctgtcggcctcgtcttgca

caaaaaggtcggcgatgcggtgaaaaaaggcgaatcgcttgtgacgatttacagcaatacggaaaatattgaagaagtcaaacaaaagcttgc

caaaagcattcgcctctcctccattcctgttgccaagccgacgcttatatacgaaaccatttcataa

TtPyNP gene sequence (AE017221.1, 1272 bp)

atgaaccccgtggtcttcatccgggagaagcgggaagggaaaaagcaccgccgggaggacctcgaggccttcctcctcggctacctgcgggac

gaggtgccggactaccaggtggccgcctggctcatggccgccttcctaaggggcctggacgccgaggagaccctctggctcaccgaaaccatgg

cccgctcggggaaggttctggacctctccggccttccccaccccgtggacaagcactcctcggggggcgtgggggacaaggtgagcctggtggtg

gggccgatcctcgccgcaagcgggtgcaccttcgccaagatgtcgggccggggcctggcccacaccggggggaccatagacaagctggagtcg

gtgccgggctggcggggggagatgacggaggcggagtttttggagagggcccggagggtgggcctcgtcatcgccgcccaaagcccggacctc

gcccccctggatgggaagctttacgccctccgcgacgtgaccgccacggtggagagcgtgcctttgatcgcgagctccatcatgagcaagaagct

cgccgccggggcgcggagcattgtcttggacgtgaaggtgggccggggggccttcatgaagaccctggaggaggcccggcttttggccaagacc

atggtggccatcggccagggggcgggaaggcgggtgagggccctcctcacctccatggaggcccccctggggcgggcggtgggcaacgccata

gaggtgcgggaggcgataggggccctcaagggggagggccccgaggatcttctggaggtggccctggccctggcggaggaggccttaaagctt

gaggggctggaccccgccctcgcccggaaggccctggagggcggggcggccttggagaagttccgggccttcctcgaggcccaggggggagac

ccccgggcggtggaggacttctcccttttgcccctcgccgaggagcaccccctccgtgccgagcgggagggcgtggtgcaggaggtggacgccta

caaggtgggcctcgccgtcctcgccctgggcggggggcggaagcggaagggggagcccattgaccacggggtgggggtctacctgctcaagaa

gcccggggaccgggtggagcggggggaggccttggccctggtctaccaccggaggcggggcctggaggaggcccttgggcacctgcgcgaggc

ctacgccctgggggaggaagcgcaccccgcccccttggtcctggaggccatctag

ApUP (full length) gene sequence (APE2105.1, 849 bp)

ttgggagacgagagtctaaggagcgccgcccgtcccgagggggagggagggctgcagtaccatctgagggtcaggaggggggatgtggcccg

ctacgttctcctcccgggagaccccgagaggacagaccttatagcccgcctctgggatgaagcgaggcttgtagcgcaccaccgggagtacagg

acgtggaccggcttctacaaggggacatcgataagtgtaacaagcaccgggataggctctcccagcacggcgatagccgttgaggagctgctga

gggttggagccgagactttcataagagtaggcactatgggcggtataagggaggatctgcggcccggcaccctggttatagggagtgcggcggt

taggatggaggggacgagcggccagtacgctccccgggggttcccagcggccgccagctatgacgttgtggcggcgctggtggaggctgctgag

gcgctcggggttaggtatgaggttggcgttgttgccagcacggacagcttctacctgggccaggggaggccggggtacggggggtatatgacgc

cggaggcttcggaagtcatacccctcctcaggtcagccggcgtcctcggcttcgagatggaggcctccgccctcttcaccctatcccagctctacg

gcgccagggcagggtgcgtgtgcgcggtagtggcaaacagggttagcggggagtttgtggtaaacgcgggggttgaagacgctgctagggttgc

ctccgaggcggtagccatactagcaggctgggacagggagagggagaagaggggtaagaaatggttttacccgagcctggcgtgcagacgca

catag

(The underlined bps represent the putative internal ribosome binding site and the bold letter mark the related potential

start codon)

Zusammenfassung

Modifizierte Nukleoside sind wertvolle Pharmazeutika, die für die Behandlung von Krebs und viralen

Erkrankungen eingesetzt werden. Außerdem dienen sie als Bausteine für die Synthese

therapeutischer Oligonukleotide mit besonderen Eigenschaften.

Während es für die Herstellung chemisch modifizierter Pyrimidinnukleoside einfache und bewährte

Verfahren gibt, stellt die organische Synthese modifizierter Purinnukleoside oft eine Herausforderung

dar, was zu mehrstufigen Verfahren mit niedriger Ausbeute führt. Die chemisch-enzymatische

Herstellung, bei der ein Pyrimidinnukleosid als Pentofuranosyl-Donor und eine Purinbase als

Pentufuranosyl-Akzeptor dient, kann daher eine attraktive Alternative sein. Für den regio- und

stereospezifischen Transfer der Zuckereinheit kommen Nukleosidphosphorylasen (NPs) als

Biokatalysatoren zum Einsatz, wobei als Substrate sowohl natürlich vorkommende, als auch

chemisch modifizierte, künstliche Vorstufen verwendet werden können. Leider ist die

Substrataktivität einer Vielzahl von hochinteressanten Nukleosid-Analoga jedoch bei den

gegenwärtig genutzten NPs sehr gering. Darüber hinaus ist es von Vorteil den Syntheseprozess bei

hohen Temperaturen durchzuführen um die Konzentration schlecht löslicher Purinbasen zu erhöhen,

doch dies führt bei vielen Enzymen zum schnellen Aktivitätsverlust. Beide Faktoren schränken den

Anwendungsbereich und die Effizienz der Synthese modifizierter Nukleoside durch NPs ein. Ziel

dieser Arbeit ist es, neue, thermostabile Varianten von NPs und ihren potentiellen Einsatz als

Biokatalysatoren zu untersuchen.

Hierfür wurden 5 NPs von 4 verschiedenen thermophilen Mikroorganismen (Deinococcus

geothermalis, Geobacillus thermoglucosidasius, Thermus thermophilus, Aeropyrum pernix) in E. coli

überexprimiert. Die Temperaturoptima und Thermostabilität der rekombinant hergestellten Enzyme

unterscheiden sich signifikant, vor allem in Abhängigkeit von Ursprungsmikroorganismus und

Enzymtyp. Die Untersuchung der Substratspezifität zeigt, dass modifizierte Nukleoside in sehr

unterschiedlichem Ausmaß als Substrate erkannt werden. Der Einsatz der vielversprechendsten

Enzym-Kombination in enzymatischen Transglykosylierungsreaktionen wurde untersucht. Hierbei

stand die Synthese 2′-fluorinierter Purinnukleoside sowie 2,6-dihalogenierte Purinnukleoside im

Fokus dieser Arbeit. 2′-Fluorinierte Nukleoside haben wertvolle pharmazeutische Eigenschaften und

verleihen synthetischen Oligonukleotiden günstige Eigenschaften, wohingegen 2,6-dihalogenierte

Purinnukleoside hervorragende Vorstufen für die Herstellung modifizierter Purinnukleoside sind. Im

Vergleich zu E. coli Enzymen, die als Biokatalysatoren für die Synthese 2′-fluorinierter

Purinnukleoside bereits in der Literatur beschrieben sind, ermöglichen die hier generierten Enzyme

die Durchführung der Transglykosylierung bei höherer Temperatur und scheinen dabei effizienter zu

sein. Gleichzeitig werden 2,6-dihalogenierte Purine sehr gut als Substrate erkannt und die Synthese

der entsprechenden (Deoxy )riboside verläuft dementsprechend schnell.

Die Ergebnisse dieser Arbeit verdeutlichen generell das Potential thermostablier Enzyme als

Biokatalysatoren und ebnen insbesondere den Weg für verbesserte und umweltfreundlichere

Verfahren für die Synthese wertvoller 2′-fluorinierter und 2,6-dihalogenierter Purinnkleosid-Analoga.

Acknowledgements

I would like to use the opportunity to thank all the people who have contributed to the

accomplishment of this work.

First of all I want to thank my supervisor Prof. Peter Neubauer for giving me the interesting research

topic, for his continuous support and for the inspiring ideas, discussions and positive view of life.

Many thanks also to Prof. Igor A. Mikhailopulo, for the intense scientific discussions, the numerous

emails from Minsk, and for sharing enthusiasm on nucleoside analogues and pictures of Murka with

me. Furthermore I would like to thank Prof. Marion Ansorge-Schumacher for helpful discussions on

enzyme technology.

Progress of this work has substantially benefitted from viable scientific collaborations and some of

the data shown have thus been obtained by other scientists in external institutions. In particular I

want to thank Olli Niemitalo and Prof. André H. Juffer for the fruitful collaboration towards the

expression optimization of DgPNP for which they have contributed the theoretical predictions of 5’

mRNA secondary structure stability and in silico sequence optimization. I am furthermore grateful for

the helpful suggestions of Thomas Schwab from the group of Prof. Reinhard Sterner, who has also

performed thermal unfolding studies with TtPyNP presented in this work. Additionally I want to

thank Knut Büttner for promptly helping out with a mass spectrometric measurement of ApMTAP

and Leif Garbe for the discussion of analytical problems. Thanks also to Marco Casteleijn for his

ongoing scientific and personal support. I want to express my gratitude to Prof. Don Cowan for giving

me the opportunity of an inspiring research stay in his lab at the other end of the world - at the

Institute for Microbial Biotechnology and Metagenomics (IMBM) of the University of the Western

Cape in Cape Town.

I am deeply thankful for the BIG-NSE scholarship of the cluster of excellence “Unifying concepts in

Catalysis” coordinated by the Technische Universität Berlin that gave me the opportunity to focus on

my research project. Special thanks to Dr. Jean-Philippe Lonjaret for his dedication to make the BIG-

NSE graduate program a success. Furthermore I would like to thank Alex Azhayev from Metkinen

Chemistry for kindly providing us with modified nucleosides for substrate screening studies.

I am grateful for the support by students with who I had the chance to work together on the

nucleoside phosphorylase project. Many thanks to Julia Geyer, who joined the project in the early

stage during her study work and with who it was always pleasant to work. Thanks to Dr. Thomas

Böhme (self proclaimed „Goldfinger“) and Dr. Bernd Janocha, whose entertaining protocols were

always enjoyable to read: Both have dared to extend their chemical expertise to the field of

biotechnology in the program „Campus Biotech“. During their 6-month practical training in our lab

they have gathered important results on ApMTAP and DgPNP expression and characterization that

are also included in this thesis.

Especially I would like to express my gratitude to Xinrui Zhou, who has joined our group in 2011,

working together with me on nucleoside phosphorylases. Her scrupulous work, her commitment and

her patience has greatly accelerated the progress of this project and the majority of data presented

in this work are the result of our productive teamwork. It has been a great pleasure to work with you.

I would like to thank the team of the department of enzyme technology, especially Andy Maraite

who helped with the first protein purification and Alexander Scholz for his invaluable work on

immobilization that is opening up future industrial perspectives.

I want to thank all the members of the IMBM in Cape Town for cordially welcoming me in- and

outside of the lab. Especially I would like to thank Heide Goodman, Ana Casanueva, Munaka Schnaka,

Randall and Layla.

Many thanks to the present and fromer co-workers at the bioprocess engineering laboratory.

Margitta Seidenstücker, Irmgard Maue-Mohn, Dirk Itzeck, and Brigtitte Burckhardt for their

unshakable drive towards a cleaner lab, their helpful advises and practical support. Herta Klein-

Leuendorf for help with bureaucracy. Stefan Junne for being ready to give competent help even the

time schedule is actually already full. My dear „kolezhanka“ Julia Glazyrina, who helped not only in

scientific matters but also with her encouraging words and pleasant coffee breaks. Jennifer Jaitzig for

sharing rooms with me, her invaluable hints and advises and for many discussions concerning science

and life. Friederike Hillig for waiting with the lunch break, Mihaela Paella and Maciek Pullarek for

their entertaining discussions, Jian Li for all the gold coins he spent. Erich for his exciting stories,

Divine for his self-propagating smile, Mirja, Eva, Kathrin and all the other co-workers that contributed

to a great working atmosphere. And the Biosilta team for sponsoring yummy sweets.

Finally I want to thank my friends, who made it easier to deal with frustrating experiments, my

parents who have supported me in every phase of my life and my brother who dragged me to do

some sports even after long working days.

Ermin – for your enormous support, your confidence and the wonderful life you share with me.