97
1 Institute of Biophysics of the CAS Department of Biophysical Chemistry and Molecular Oncology Monika Hermanová Electrochemical methods for detection of DNA- protein interactions and for monitoring of DNA enzymatic processing Ph.D. Dissertation Supervisor: Doc. RNDr. Miroslav Fojta, CSc. Brno 2018

protein interactions and for monitoring of DNA enzymatic

Embed Size (px)

Citation preview

1

Institute of Biophysics of the CAS Department of Biophysical Chemistry and Molecular Oncology

Monika Hermanová

Electrochemical methods for detection of DNA-

protein interactions and for monitoring of DNA

enzymatic processing

Ph.D. Dissertation

Supervisor: Doc. RNDr. Miroslav Fojta, CSc. Brno 2018

2

Bibliographic entry

Author: Mgr. Monika Hermanová

Title of Dissertation: Electrochemical methods for detection of DNA-protein interactions

and for monitoring of DNA enzymatic processing

Degree Program: Biology

Field of Study: Molecular and cell biology

Supervisor: doc. RNDr. Miroslav Fojta, CSc.

Academic Year: 2017/2018

Number of Pages: 128

Key words: Electrochemistry, DNA-protein interactions, p53 protein, DNA labeling,

terminal deoxynucleotidyl transferase

3

Bibliografický záznam

Jméno a příjmení autora: Mgr. Monika Hermanová

Název disertační práce: Elektrochemické metody pro detekci interakcí DNA s proteiny a pro

sledování enzymatických přeměn DNA

Název disertační práce anglicky: Electrochemical methods for detection of DNA-protein

interactions and for monitoring of DNA enzymatic processing

Studijní program: Biologie

Studijní obor: Molekulární a buněčná biologie

Školitel: doc. RNDr. Miroslav Fojta, CSc.

Akademický rok: 2017/2018

Počet stran: 128

Klíčová slova v češtině: Elektrochemie, DNA-protein interakce, protein p53, značení DNA,

terminální deoxynukleotidyl transferáza

Klíčová slova v angličtině: Electrochemistry, DNA-protein interactions, p53 protein, DNA

labeling, terminal deoxynucleotidyl transferase

4

© Monika Hermanová, Masaryk University, 2018

5

I would like to thank my supervisor, doc. RNDr. Miroslav Fojta, CSc. for all the

support that I received from him during my Ph.D. studies, for a lot of useful advice and ideas

and for being such a scientist I would like to become one day.

My thanks also go to all current and former colleagues at the Department of

Biophysical Chemistry and Molecular Oncology, especially Hanka Pivoňková, Peťa Orság,

Pavlínka Havranová, Luděk Havran and Honza Špaček for being extremely helpful at any

occasion I needed something and for creating a wonderful working environment where it has

always been fun to work.

I would also like to thank my whole family because without them, this work would

never come into existence. My mom and dad for being very supportive during my childhood

and my studies, my husband Michal for helping me with anything that I did not manage to do

and all of them - my parents, Michal and my mother-in-law for all the babysitting they did.

And finally, my biggest thanks of all go to Toník, who had a great patience with his

mom when she was too busy to play with him and who was often the last kid in the

kindergarten in the afternoon because mom had to finish her experiments in the lab. Despite

all of this, he has always been in a good mood and has helped me handle everything with

ease.

6

Abstract

Three novel applications of DNA electrochemistry and redox DNA labeling are

presented. Electrochemical signals of DNA are based either on redox activity of DNA bases,

which can be observed as CA and G peak at mercury electrodes and G and A peak at carbon

electrodes or on adsorption/desorption and reorientation processes which can be observed as

tensammetric peaks at mercury electrodes. Another option for studying DNA using

electrochemical methods lies in its redox labeling, which can be achieved using covalently

bound labels, such as osmium tetroxide complexes, or by enzymatic incorporation of

modified nucleobases into DNA.

Redox labeling of DNA via enzymatic incorporation was used in development of a

novel approach for detection of DNA-protein interactions. This approach is based on dual

labeling of DNA probes using two different electroactive labels, which yield reduction peaks

at distinct potentials and therefore enable detection of protein binding in a competition

experiment. We show that using this approach, specific and non-specific binding of the p53

protein can be distinguished as a strong preference of the p53 protein was observed towards

DNA probes bearing a specific p53 binding site (p53CON).

Label-free detection using intrinsic DNA signals was used in studies of terminal

deoxynucleotidyl transferase (TdT) tailing reactions. We found out that electrochemical

detection can be very useful for monitoring the TdT tailing reactions, especially with

pyrolytic graphite electrode (PGE) being suitable for remarkably precise determination of the

tailing reaction products length. On the other hand, hanging mercury drop electrode (HMDE)

revealed formation of various DNA structures, such as DNA hairpins and G-quadruplexes,

which influence behavior of DNA molecules at the negatively charged surface of HMDE as

well as progress of the TdT tailing reaction.

Modification with osmium tetroxide complex was applied in development of a new

two-step technique of DNA modification, which comprised enzymatic construction of DNA

bearing butyl acrylate (BA) moieties followed by chemical modification of a reactive C=C

double bond in the acrylate residue. Such approach enabled modification of the BA

conjugates in both single- and double-stranded (ds) DNA under conditions precluding

modification of nucleobase residues in ds DNA.

7

Abstrakt

V této práci představujeme tři nové aplikace využívající elektrochemii DNA a

redoxní značení DNA. Elektrochemické signály, které poskytuje DNA, jsou založeny buď na

redoxní aktivitě bází DNA, v důsledku které můžeme pozorovat píky CA a G na rtuťových

elektrodách a píky G a A na uhlíkových elektrodách, nebo na adsorpčně-desorpčních jevech a

reorientaci DNA na povrchu elektrody, které můžeme sledovat prostřednictvím

tenzametrických píků na rtuťových elektrodách. Další možnost pro využití elektrochemických

metod ke studiu DNA představuje redoxní značení DNA, kterého může být dosaženo

například pomocí kovalentně se vázajících značek, jako jsou komplexy oxidu osmičelého,

nebo pomocí enzymatické inkorporace modifikovaných nukleobází do DNA.

Značení DNA prostřednictvím enzymatické inkorporace redoxních značek do DNA

bylo využito v nové metodě pro detekci interakcí proteinů s DNA. Tato metoda je založena na

duálním značení DNA sond s využitím dvou různých elektroaktivních značek, které poskytují

redukční píky při odlišných potenciálech a umožňují tak detekci vazby proteinu na značené

sondy v kompetičním upořádání. S použitím tohoto přístupu bylo možné rozlišit sekvenčně

specifickou a nespecifickou vazbu proteinu p53 na DNA, jelikož protein p53 vykazoval silnou

preferenci pro sondy obsahující specifickou vazebnou sekvenci pro protein p53 (p53CON).

Detekce DNA pomocí jejích vlastních signálů byla využita pro studium syntézy

jednořetězcových úseků DNA katalyzované terminální deoxynukleotidyl transferázou (TdT).

Zjistili jsme, že elektrochemická detekce může být úspěšně použita pro sledování této reakce.

Obzvláště elektroda z pyrolytického grafitu se ukázala jako velmi vhodná pro pozoruhodně

přesné určení délky produktů reakce. Na druhou stranu, visící rtuťová kapková elektroda

(HMDE) odhalila tvorbu různých struktur DNA, jako jsou vlásenky nebo G-kvadruplexy,

které ovlivňují chování molekul DNA na negativně nabitém povrchu HMDE a také průběh

syntézy DNA katalyzované TdT.

Modifikace s využitím komplexu oxidu osmičelého byla uplatněna v nové

dvoukrokové technice modifikace DNA, která spočívala v enzymatické konstrukci DNA

nesoucí butylakrylátové (BA) skupiny a následné modifikaci reaktivní dvojné vazby C=C

v akrylátovém zbytku. Tento přístup umožnil modifikaci BA jak v jednořetězcové, tak

dvouřetězcové DNA i za podmínek, za kterých modifikace bází v dvouřetězcové DNA

neprobíhá.

8

Contents

1. INTRODUCTION .............................................................................................................. 10

2. PROTEIN-NUCLEIC ACIDS INTERACTIONS ........................................................... 10

2.1 Proteins interacting with nucleic acids ........................................................................... 10

2.2 p53 protein ........................................................................................................................ 11

2.2.1 p53 function ................................................................................................................. 12

2.2.2 p53 structure ................................................................................................................ 13

2.2.3 p53-DNA interactions .................................................................................................. 15

2.3 Terminal Deoxynucleotidyl Transferase ........................................................................ 16

2.4 Methods for studying protein-nucleic acids interactions .............................................. 19

2.4.1 Electrophoretic mobility shift assay ............................................................................ 20

2.4.2 Immunoprecipitation techniques .................................................................................. 21

2.4.3 DNA footprinting ......................................................................................................... 21

2.4.4 Other methods .............................................................................................................. 21

2.4.5 Electrochemical techniques ......................................................................................... 21

3. ELECTROANALYTICAL CHEMISTRY ...................................................................... 22

3.1 Types of electrodes ........................................................................................................... 23

3.1.1 Mercury electrodes ...................................................................................................... 24

3.1.2 Solid electrodes ............................................................................................................ 25

3.2 Electrochemical methods ................................................................................................. 26

3.2.1 Cyclic Voltammetry (CV) ........................................................................................... 27

3.2.2 Linear Sweep Voltammetry (LSV) .............................................................................. 27

3.2.3 Differential-pulse voltammetry (DPV) ........................................................................ 28

3.2.4 Square-wave voltammetry (SWV) ............................................................................... 28

3.2.5 Alternating current voltammetry (ACV) ..................................................................... 29

3.3 Electrochemistry of nucleic acids .................................................................................... 29

3.3.1 Adsorptive transfer stripping in DNA analysis ............................................................ 30

9

3.3.2 Analysis of DNA structure at mercury electrodes ....................................................... 31

3.3.3 DNA signals at mercury electrodes ............................................................................. 31

3.3.4 DNA signals at carbon electrodes ................................................................................ 33

3.4 Electrochemical labeling of nucleic acids ....................................................................... 34

3.4.1 Noncovalently bound redox indicators ........................................................................ 35

3.4.2 Osmium tetroxide complexes ...................................................................................... 35

3.4.3 Redox labels and their enzymatic incorporation .......................................................... 37

3. AIMS OF THE DISSERTATION .................................................................................... 40

4. RESULTS AND DISCUSSION ......................................................................................... 41

4.2 Label-free voltammetric detection of products of terminal deoxynucleotidyl

transferase tailing reaction .................................................................................................... 41

4.3 Butylacrylate-nucleobase Conjugates as Targets for Two-step Redox Labeling of

DNA with an Osmium Tetroxide Complex .......................................................................... 47

5. CONCLUSIONS ................................................................................................................. 51

6. REFERENCES ................................................................................................................... 53

LIST OF ABBREVIATIONS ................................................................................................ 67

LIST OF PUBLICATIONS AND CONFERENCES .......................................................... 70

APPENDICES ......................................................................................................................... 72

APPENDIX 2 .......................................................................................................................... 73

APPENDIX 3 .......................................................................................................................... 91

10

1. Introduction

Electrochemistry of nucleic acids is a booming field of study, although its beginning

dates back to 1950s when the first electrochemical signals of DNA were discovered by Emil

Paleček at the Institute of Biophysics in Brno. Since then, many applications have been

introduced which take advantage of either intrinsic electrochemical activity of DNA bases at

various electrodes or chemical modifications of DNA. These applications involve

development of DNA hybridization sensors, approaches enabling studies of DNA structure

and damage or detection of DNA-protein interactions.

Proteins including enzymes interacting with DNA are of crucial importance as they

influence and regulate many key cellular processes therefore new methodologies for studying

such interactions are required. We can monitor either the interaction itself (like in the case of

p53 protein interactions with DNA presented in this work) or results of the enzymatic

processing of DNA (like in the case of terminal deoxynucleotidyl transferase tailing reaction

also presented in this work). These two approaches presented here are based on

electrochemistry of nucleic acids, with the first one utilizing redox DNA labeling and the

second one taking advantage of label-free detection using intrinsic signals provided by DNA.

A novel approach for labeling DNA based on modification with osmium tetroxide complexes

also presented in this work can be potentially applied for similar applications in the future.

2. Protein-nucleic acids interactions

2.1 Proteins interacting with nucleic acids

Interactions of proteins with nucleic acids (NA) are of particular significance since

they occur in the most important processes in the cell, such as replication, transcription or

DNA repair. Proteins can interact with either RNA or DNA during these processes,

nevertheless, in this work, only DNA is subject of the study. Protein–DNA interactions can

vary a lot in their nature: some proteins are able to bind non-specifically to any DNA

sequence while other proteins are only able to bind to precisely defined genomic regions.

Sequence-specific DNA binding is typical for transcription factors or restriction

nucleases – proteins which need to locate the precise sequence in the genome in order to

perform its function. Nevertheless, slight base pair alterations in the DNA can be often

overcome. During sequence-specific DNA binding, protein interactions with NA sequences is

11

primarily determined by recognition of hydrogen-bonding determinants situated in the major

and minor grooves of the DNA that interact with complementary recognition of the amino

acids of the protein itself (Seeman et al., 1976). Hence, interactions are promoted when DNA

bases are arranged in an optimal sequence. The hydrogen bond donator and acceptor patterns

in the DNA grooves are recognized by complementary hydrogen bond donator and acceptors

on the protein surface (Bowater et al., 2015).

The most frequent mode of DNA binding by proteins is the non-specific DNA

binding. The main component providing the necessary free energy to stabilize the interaction

is the electrostatic attraction between positively charged amino acids and the negatively

charged phosphodiester DNA backbone (Jen-Jacobson, 1997). These electrostatic, non-

specific binding affinities are based on the displacement of counter-ions from the DNA

(Hippel, 1994) and, thus, are not as tight as the previously described sequence-specific

interactions. Although the proteins binding DNA in a non-specific manner do not require a

specific sequence, they may prefer or demand specific DNA structures. Such DNA regions

can be essential, specialized sites that require a non-B structure, some arise accidently during

various cellular processes, and others can be damage induced (Bowater et al., 2015).

There is a wide range of proteins that can bind DNA, including histones, proteins

involved in replication or transcription such as helicases, single-stranded DNA binding

proteins or transcription factors, proteins involved in DNA repair and various enzymes acting

on DNA such as DNA polymerases, nucleases, ligases or topoisomerases. In this work, two

proteins belonging to different classes of proteins capable of DNA binding were used – one of

them, the p53 protein, is a transcription factor, the other one, terminal deoxynucleotidyl

transferase, is a special type of a DNA polymerase. Both of them are therefore described in

more detail.

2.2 p53 protein

The p53 protein, also called the “guardian of the genome” (Lane, 1992), is a tumor

suppressor which has been extensively studied over the past couple of decades. It was first

described in 1979 as a protein that binds to the simian virus (SV40) large T antigen (Lane and

Crawford, 1979). Its importance for maintaining genetic stability of a cell and preventing

cancer transformation is evident from the fact that more than 50 % of human cancers contain

mutations in this gene (Levine, 1997). p53 functions largely as a transcription factor, and can

trigger a variety of antiproliferative programs by activating or repressing key effector genes

(Zilfou and Lowe, 2009), which can act for example in cell cycle, DNA repair or apoptosis.

12

The biological activities of p53 are closely connected with its ability to interact with DNA

(Fojta et al., 2004).

There are two structural and functional homologs of the p53 protein, p63 and p73. As

a result of sharing similar domain architecture and sequence identity with p53, p73 and p63

isoforms can form oligomers, bind DNA and transactivate majority of the p53-responsive

genes (Dötsch et al., 2010; Khoury and Bourdon, 2010).

2.2.1 p53 function

P53 is a short-lived protein that is maintained at low, often undetectable, levels in

normal cells. Stabilization of the protein in response to a stress signal (e.g. DNA damage)

results in a rapid rise in p53 levels (Bates and Vousden, 1996; Kubbutat et al., 1997) and

subsequent activation of target genes. Changes in the rate of transcription of the p53 gene thus

play a minor role, if any, in induction of p53 (Oren, 1999). One of the key mechanisms by

which the p53 function is controlled, is through interaction with MDM2 protein (Kubbutat et

al., 1997). MDM2 protein acts as E3 ubiquitin ligase, which enables it to ubiquitinate the p53

protein and target it for degradation in proteasome (Michael and Oren, 2003). The p53 protein

positively regulates the MDM2 gene at the level of transcription and the MDM2 protein

negatively regulates the p53 protein at the level of its activity. This creates a feedback loop

that regulates both the activity of the p53 protein and the expression of the MDM2 gene (Wu

et al., 1993).

After p53 activation, several outcomes are possible, depending on type and extent of

the stress factors affecting the cell, all of which may under appropriate circumstances

contribute to its tumor suppressive properties. In the case of DNA damage, p53 can arrest cell

cycle in G1 or G2 phase in order to allow DNA damage repair occur prior to replication of

DNA or beginning of mitosis. p53 induces a G1 arrest primarily through the transactivation of

p21Waf1/Cip1

, a cyclin-dependent kinase inhibitor (Brugarolas et al., 1995; Deng et al., 1995),

which causes activation of Rb protein, resulting in negative regulation of the E2F family of

transcription factors and blockage of the transfer from G1 to S phase (Giono and Manfredi,

2006).

After the induction of cell cycle arrest, p53 can activate genes involved in DNA repair

or can directly interact with proteins taking part in the DNA repair process (Latonen and

Laiho, 2005). One of the most studied proteins, which take part in the DNA repair and are

transactivated by p53, is the GADD45 protein (Smith et al., 2000). If the DNA damage

13

cannot be repaired, p53 can induce p53-dependent apoptosis. There are three mechanisms of

apoptosis induced by p53: binding of a ligand to death receptors from the TNF-R (tumor

necrosis factor receptor) family, a mechanism involving translocation of p53 to mitochondrial

membrane and induction of apoptosis in a transcription-independent manner by p53 cytosolic

activities (Bourdon et al., 2002; Wang et al., 2005; Green and Kroemer, 2009). Several p53

target genes including Bax, PUMA, Noxa, p53-AIP, PIG3, Fas/APO1, and KILLER/DR5

have been implicated in p53-induced apoptosis (Owen-Schaub et al., 1995; Toshiyuki and

Reed, 1995; Polyak et al., 1997; Oda, 2000).

In the case that the p53 protein is mutated, the cell processes in which p53 is involved

can be altered by different mechanisms: a) loss of the wtp53 function, b) dominantly negative

inhibition of the wtp53 function and c) gain of function, which can lead to transcriptional

activation of genes involved in cell proliferation, cell survival and angiogenesis.

Consequently, cells expressing some forms of mutant p53 show enhanced tumorigenic

potential with increased resistance to chemotherapy and radiation (Cadwell and Zambetti,

2001).

2.2.2 p53 structure

Human p53 is a nuclear phosphoprotein of MW 53 kDa, encoded by a 20-Kb gene

containing 11 exons and 10 introns, which is located on the small arm of chromosome 17

(Lamb and Crawford, 1986). Wild-type p53 protein contains 393 amino acids and is

composed of several structural and functional domains (Fig. 1): a N-terminus containing an

amino-terminal domain (residues 1-42) and a proline-rich region with multiple copies of the

PXXP sequence (residues 61-94, where X is any amino acid), a central core domain (residues

102-292), and a C- terminal region (residues 301-393) containing an oligomerization domain

(residues 324-355), a strongly basic carboxyl-terminal regulatory domain (residues 363-393),

a nuclear localization signal sequence and a nuclear export signal sequence (Bai and Zhu,

2006).

14

Figure 1. Structure of the p53 protein. The protein has three main domains: N-terminal

activation domain, DNA binding domain and C-terminal tetramerization domain.The N-

terminal domain includes transactivation subdomain and a proline-rich fragment. The central

DNA binding domain is required for sequence-specific DNA binding. The C-terminal region

is considered to perform a regulatory function. Numbers indicate residue number. NLS,

nuclear localization signal sequence; NES, nuclear export signal sequence. (Bai and Zhu,

2006)

The N-terminal domain consists of two transactivation domains and a proline-rich

region. It is required for transactivation activity and interacts with various transcription factors

including acetyltransferases and MDM2 (Fields and Jang, 1990). The proline-rich region

plays a role in p53-dependent apoptosis and in p53 stability regulated by MDM2 (Sakamuro

et al., 1997).

Central domain is responsible for the sequence-specific DNA binding to the consensus

binding site. Amino acid residues within this domain are frequently mutated in human cancer

cells and tumor tissues. The Arg175, Gly245, Arg248, Arg249, Arg273, and Arg282 (see Fig.

X) are reported to be mutation hot spots in various human cancers (Bai and Zhu, 2006). The

central region of p53 is its most highly conserved region, not only when p53 is compared with

its homologues from Drosophila and Caenorhabditis elegans, but also as compared with its

mammalian family members, p63 and p73 (Kaelin, 1999).

C-terminus of the p53 protein contains an oligomerization domain, which is located

within the amino acids 324 and 355 and is responsible for tetramerization of the protein and

the C-terminal regulatory and DNA-binding domain. This C-terminal domain of p53 can act

as a negative regulator of the core DNA binding domain (Hupp, 1999). If the interaction

between the C-terminus and the core DNA binding domain is disrupted by posttranslational

modification (such as phosphorylation and acetylation), the DNA binding domain will

become active, thus induce an enhanced transcriptional activity (Bai and Zhu, 2006). The C-

terminal domain can bind DNA in a sequence-nonspecific manner.

15

2.2.3 p53-DNA interactions

As has been mentioned above, p53 acts as a transcription factor and therefore its

function is closely related to its ability to bind DNA. p53 can bind DNA either in a sequence-

specific manner, which is mediated by the central DNA binding domain, or in a sequence-

nonspecific (structure-selective) manner, mediated by the C-terminal DNA binding domain.

2.2.3.1 Sequence-specific DNA binding

For binding DNA in a sequence-specific manner, p53 protein requires a consensus

binding site consisting of two copies of the 10 base pair motif (p53CON) 5‘-

PuPuPuC(A/T)(T/A)GPyPyPy-3‘ separated by 0-13 base pairs (El-Deiry et al., 1992), which

is found in the vicinity of many of its target gene promoters (McKinney and Prives, 2002).

The internal symmetry of the motif and the fact that the consensus binding site contains two

copies of the motif imply that p53 binds DNA as a tetramer. One copy of the motif is

insufficient for the binding, and subtle alterations of the motif, even when present in multiple

copies, result in loss of affinity for p53 (El-Deiry et al., 1992). Sequences, that match the

consensus binding site, are often bound with different efficacy, on the other hand, some

sequences, which do not correspond to the consensus site, are bound with high affinity (Foord

et al., 1993; Halazonetis et al., 1993). Intensity of the p53 binding also depends on number of

nucleotides which are inserted between the two copies of the motif. While insertion of 5 or 15

nucleotides inhibits p53 binding to DNA, insertion of 10 nucleotides does not affect the

binding, which corresponds to the fact that after insertion of 10 nucleotides, the half-sites

remain on the same face of the double helix (Wang et al., 1995). DNA topology is also an

important parameter for regulating the specific interaction of p53 with its target binding sites

(Göhler et al., 2002).

Aside from the central DNA binding site, the C-terminal domain of the p53 protein is

also very important for the sequence-specific DNA binding as it regulates the sequence-

specific binding of the central DNA binding domain. If not modified, the C-terminal domain

inhibits the central DNA binding domain but after phosphorylation of serines or acetylation of

lysines in the C-terminal domain, the sequence-specific DNA binding is activated (Hupp and

Lane, 1994), which is caused by the diminution of the negative regulatory effect of the C-

terminal domain. Same effect can be achieved by antibodies, peptides or other molecules

interacting with the C-terminal domain (Selivanova et al., 1998) or by deletion of the last 30

amino acids (Hupp and Lane, 1994).

16

2.2.3.2 Sequence-nonspecific DNA binding

Besides the sequence-specific DNA binding, p53 is able to bind also DNA not

containing the consensus binding site. It was found that p53 binds selectively to supercoiled

DNA, which has been denominated as supercoil-selective (SCS) DNA binding (Paleček et al.,

1997; Palecek et al., 2001). Negatively supercoiled (sc) DNA, regardless of the presence or

absence of the p53CON, is bound preferentially by both wt (Paleček et al., 1997; Mazur et al.,

1999; Fojta et al., 2004) and mutant (Brázdová et al., 2009) p53. Basic segment of the C-

terminal DNA binding domain (amino acids 363–382) in an oligomeric state is responsible

for the SCS binding (Brázdová et al., 2002; Fojta et al., 2004).

The C-terminal DNA binding domain can recognize non-canonical DNA structures

stabilized by supercoiling, such as hairpins (Palecek et al., 2001; Fojta et al., 2004), cruciform

structures (Jagelská et al., 2010), bent DNA (McKinney and Prives, 2002), three- and four-

way junctions (Subramanian and Griffith, 2005), structurally flexible chromatin DNA (Kim

and Deppert, 2003), telomeric t-loops (Stansel et al., 2002) or triplexes (Brázdová et al.,

2016) and bind them with substantial preference.

The sequence-nonspecific DNA binding comprises p53 binding to DNA exhibiting

various types of damage. This can include single-stranded DNA ends (Bakalkin et al., 1995),

DNA with insertion-deletion mismatches, DNA containing single- or double-strand breaks

resulting from UV or ionizing radiation or DNA modified with anticancer drugs, such as

cisplatin (Fojta et al., 2003; Pivoňková et al., 2006).

2.3 Terminal Deoxynucleotidyl Transferase

Terminal Deoxynucleotidyl transferase (TdT) is a DNA polymerase, which is unique among

other polymerases as it possesses an unusual ability to incorporate nucleotides in a template-

independent manner using only single-stranded DNA as the nucleic acid substrate (Michelson

and Orkin, 1982). TdT was first isolated from a calf thymus gland and was described as

enzyme with polymerase activity (Bollum, 1960). Although the biochemical mechanism of

TdT reaction was described soon (Kato et al., 1967), physiological role of TdT remained

unclear for many years. However, later it was discovered that its physiological role lies in

random addition of nucleotides to single-stranded DNA during V(D)J recombination

(Baltimore, 1974). During the assembly of immunoglobulin and T cell receptor variable

region genes from variable (V), diversity (D), and joining (J) segments, the germline-encoded

repertoire is further diversified by processes that include template-independent addition of

nucleotides (N regions) by TdT at gene segment junctions. TdT deficient lymphocytes have

17

no N regions in their variable region genes, which shows that TdT is responsible for N region

addition (Komori et al., 1993). The ability of TdT to randomly incorporate nucleotides

increases antigen receptor diversity and aids in generating the ∼1014

different

immunoglobulins and ∼1018

unique T cell antigen receptors that are required for the

neutralization of potential antigens (Sadofsky, 2001; Motea and Berdis, 2010). TdT is thus

crucial for evolution and adaptation of the vertebrate immune system (Kunkel et al., 1986;

Komori et al., 1993).

For its catalytic activity, the enzyme requires a single-stranded initiator that is at least three

nucleotides long and a free 3'-OH end for extension (Kato et al., 1967; Bollum, 1974). The

order by which TdT binds DNA and dNTP is random; studies performed in the absence or

presence of product inhibitors suggest a rapid equilibrium random mechanism in which TdT

forms the catalytic competent ternary complex via binding of dNTP prior to DNA or vice

versa (Deibel and Coleman, 1980). This also implies that TdT catalyzes DNA synthesis in a

distributive mode. Single-stranded DNA is generally the preferred primer for the reaction but

under certain conditions, TdT can add tails to double-stranded DNA as well (Roychoudhury

et al., 1976; Deng and Wu, 1981). The enzyme also catalyzes a limited polymerization of

ribonucleotides at the 3'-end of oligodeoxynucleotides (Roychoudhury et al., 1976; Boulé et

al., 2001).

Since the DNA synthesis by TdT is template-independent, it could be expected that TdT

incorporates all natural nucleotides with equal efficiencies. Nevertheless, although in vitro the

TdT can incorporate all four nucleotides onto single-stranded DNA, in vivo, a bias was

observed for the incorporation of dGMP and dCMP versus dAMP and dTMP (Basu et al.,

1983; Mickelsen et al., 1999). This preference may offer explanation for high G/C content of

the Ig and T cell receptor N regions. For example, it has been demonstrated that Km for dGTP

and dATP can differ very significantly, with the Km for dGTP being more than 4-fold lower

than Km for dATP (Modak, 1978; Yang et al., 1994). At the molecular level, the preference in

nucleotide utilization could reflect favorable hydrogen-bonding interactions between the

incoming dNTP and active site amino acids that guide nucleotide binding (Motea and Berdis,

2010).

Another unique feature attributed to TdT, aside from the template-independent activity, is the

ability to perform de novo synthesis of short fragments of DNA ranging in size from 2- to 15-

mers when provided with dNTPs in the absence of a primer (Ramadan et al., 2004). These

fragments are hypothesized to act as signals for DNA repair or recombination machinery

18

(Ramadan et al., 2004). However, this hypothesis has yet to be tested, in order to prove if de

novo DNA synthesis indeed occurs in vivo and to show biological relevance of this

phenomenon.

TdT, like all DNA polymerases, requires divalent metal ions to catalyze the phosphoryl

transfer reaction associated with nucleotide incorporation. However, TdT is unique in its

ability to use a variety of divalent cations such as Co2+

, Mn2+

, Zn2+

and Mg2+

(Deibel and

Coleman, 1980). Each metal ion has different effects on the kinetics of nucleotide

incorporation. For example, Mg2+

facilitates the preferential utilization of dGTP and dATP

whereas Co2+

increases the catalytic polymerization efficiency of the pyrimidines, dCTP and

dTTP (Chang and Bollum, 1990). Zn2+

behaves as a unique positive effector for TdT since

reactions with Mg2+

are stimulated by addition of micromolar quantities of Zn2+

. However,

Zn2+

is not an intrinsic part of the enzyme (Chang and Bollum, 1990). X-ray structures and

modeling of the 3’-end during the pre-catalytic and post-catalytic state for aforementioned

divalent cations are depicted in Fig. 2.

19

Figure 2. Pre- and post-catalytic active sites with different metal ions (Gouge et al., 2013).

TdT is able to tolerate various bulky modifications of the nucleotides. This feature has been

utilized in in vivo and in vitro labeling of double-strand DNA breaks in a technique called

TUNEL (TdT-mediated dUTP-biotin nick end-labeling) (Motea and Berdis, 2010). The

nucleotide analogs, which can be incorporated by TdT, include 2′,3′-dideoxynucleotides

(Ono, 1990), p-nitrophenylethyl triphosphate or p-nitrophenyl triphosphate (Arzumanov et

al., 2000) and recently in our lab it was shown that TdT can incorporate 3-nitrophenyl-7-

deazaG (Horáková et al., 2011), butyl acrylate modified dATP or dUTP (Havranová-

Vidláková et al., 2017) or vinyl (unpublished results).

2.4 Methods for studying protein-nucleic acids interactions

As already mentioned, protein-nucleic acids interactions are of crucial importance in

many key cellular processes, which leads to the need for extensive studies of these

20

interactions. Many traditional methods are used for this purpose along with new methods,

which are being developed.

Each technique can provide a wealth of knowledge of such interactions, but each has

limitations that usually restrict it from elucidating a full description of the mode of interaction

between the protein and NA. Hence, alternative, complementary techniques are usually

applied to the same system to provide a more complete description of these interactions

(Bowater et al., 2015).

2.4.1 Electrophoretic mobility shift assay

The electrophoretic mobility shift assay (EMSA) is based on electrophoretic

separation of protein–NA mixtures. The speed, at which different molecules move through the

gel, is determined by their size, charge and shape. After the protein binding to the NA, the

complex of NA bound to protein becomes less mobile and is shifted up the gel compared to

the NA alone. The fraction of free and bound NA molecules can thus be determined from the

ratio of bound and unbound NA.

Competition assay is a type of an EMSA experiment, which enables to determine the

most favorable DNA sequence for the binding protein. Different oligonucleotides of defined

sequence are used as competitors. Choice of appropriate competitors allows identification of

the precise binding site of the protein. Protein binding to different DNAs can be also

indirectly evaluated using indicator oligonucleotide or longer DNA substrate as competitor

(Brázda et al., 2006).

Another type of the EMSA experiment comprises use of antibody that recognizes the

protein. When the antibody is added to the mixture, larger complex containing the NA,

protein and antibody is created and is even more shifted. This method is referred to as a

supershift assay and is used to unambiguously identify a protein present in the protein–NA

complex (Bowater et al., 2015).

EMSA analysis is quite limited by the fact that experimental environments are

restricted by the conditions required for electrophoresis, which constitutes quite a

disadvantage of this method. The ionic strength of a solution has a substantial impact on the

modes of protein–NA interactions, meaning non-specific interactions that generally rely

heavily on charge are reduced in high-salt environments, but stronger specific interactions that

are dependent on other factors, such as base sequence, can remain (Jen-Jacobson, 1997). This

21

can result in misinterpretation of the observed data when DNA binding is studied using

EMSA analysis.

2.4.2 Immunoprecipitation techniques

In immunoprecipitation techniques, antibodies are used to pull down their antigen out

of solution; the antigen can be either the protein binding to NA or a specific sequence or

structure in the NA. To allow recovery and analysis of the sample after the precipitation, the

agent that recognizes the complex is often immobilized to a substrate or surface. An important

advantage of precipitation approaches is their significant degree of flexibility in terms of how

the experiment can be set up, allowing for diverse reaction conditions to be studied.

Chromatin immunoprecipitation (ChIP) is a widespread method for identification of

sequences to which proteins bind in genomic DNA within cells (Furey, 2012; Christova,

2013).

2.4.3 DNA footprinting

Footprinting assays are based on the principle of protection of protein-bound DNA

from degradation. The procedure employs chemical (using e.g. hydroxyl radicals) or

enzymatic (using e.g. DNase I) digestion of naked- and protein bound-DNA oligomers. Both

the reactions are then compared using gel electrophoresis. Footprinting has been a valuable

tool for elucidating sequence specificity and dissociation constants of a variety of ligands

binding to DNA (Dey et al., 2012). Its advantage lies in the fact that the protein-DNA binding

can occur under defined reaction conditions.

2.4.4 Other methods

Many other methods have been employed in studies of protein-nucleic acids

interactions. These include high-resolution techniques such as X-ray crystallography or

nuclear magnetic resonance (Bowater et al., 2015). Further examples of techniques used for

studying protein-NA interactions are isothermal titration calorimetry (Bowater et al., 2015),

yeast one-hybrid assay, phage display for DNA-binding proteins, proximity ligation assay,

fluorescence resonance energy transfer (FRET), circular dichroism, atomic force microscopy

(AFM), surface plasmon resonance (SPR) or various in silico tools for prediction and

identification of DNA–protein interactions (Dey et al., 2012).

2.4.5 Electrochemical techniques

Among the widely used biophysical methods based on optical detection, new methods

utilizing electrochemical detection have been presented. They can take advantage of the

22

intrinsic electrochemical or electrocatalytic activity of either proteins or DNA. Approaches

utilizing signals specific for proteins enable detection of their DNA binding, for example

lysozyme binding to DNA aptamers (Kawde et al., 2005; Ostatná et al., 2017) or MutS

protein recognizing mispaired and unpaired bases in duplex DNA (Paleček et al., 2004;

Masařík et al., 2007). Other approaches are based on detection of DNA, both natural and

labeled. DNA can be immobilized on the electrode surface, which has been used in the study

of T3-RNA polymerase (Meunier-Prest et al., 2010), protein MutH (Ban et al., 2004), anti-

DNA antibodies (Evtugyn et al., 2008), -thrombin (Evtugyn et al., 2008; Ding et al., 2010)

or lysozyme (Huang et al., 2009) binding to DNA. Another approach utilizes separation of

protein-DNA complexes at magnetic beads (MB) (Paleček and Fojta, 2007); subsequent

electrochemical detection can be label-free, where p53 protein binding to supercoiled and

linearized plasmid DNA can be distinguished (Němcová et al., 2010) or can use DNA

labeling by various methods such as terminal deoxynucleotidyl transferase tail-labeling

(Horáková et al., 2011) or labeling by oxoosmium complexes (Němcová et al., 2014).

Magnetic beads-based assay was also used for detection of recognition of a specific aptamer

by thrombin (Cheng et al., 2010; Zheng et al., 2010). Another possibility for detection of

protein-DNA binding lies in use of redox-labelled click-transformable DNA probe, which has

been applied in a study using click reaction of azidophenyl for detection of p53-DNA binding

(Balintová et al., 2015). Electrochemical detection can be also employed in studying damaged

DNA and enzymes involved in its processing (Fojta et al., 2016) such as T4 endonuclease V

and E.Coli exonuclease III (Fojta, 2005; Havran et al., 2008), 8-oxoguanine DNA glycosylase

(Liu et al., 2015), photolyase (DeRosa et al., 2005) or ligase (Zauner et al., 2005; Vacek et

al., 2008).

3. Electroanalytical chemistry

Electroanalytical chemistry is the branch of chemistry concerned with the interrelation

of electrical and chemical effects. A large part of this field deals with the study of chemical

changes caused by the passage of an electric current and the production of electrical energy by

chemical reactions. In electrochemical systems, we are concerned with the processes and

factors that affect the transport of charge across the interface between chemical phases, which

are an electronic conductor (an electrode) and an ionic conductor (an electrolyte). Since one

interface is not enough to perform the measurement, electrochemical experiments take place

in collections of interfaces called electrochemical cells. These systems comprise at least two

23

electrodes separated by at least one electrolyte phase. Instead of the two-electrode system

with a working and reference electrode, often a three-electrode system is used, which is

composed of working electrode, counter (auxiliary) electrode and a reference electrode. In

this system current flows and is measured between working and counter (or auxiliary)

electrode, while the potential is measured (and controlled) with respect to the reference

electrode (Fig. 3).

Figure 3. Schematic view of a three electrode system, with working, reference and

counter electrode.

Electrochemical methods enable to study electroactive chemical species or species

exhibiting surface activity. Electroactive species can be studied owing to faradaic currents,

which involve electron transfer across the electrode-electrolyte interface. This feature enables

detection of reduction or oxidation reactions occurring when a certain potential is applied.

During reduction the electroactive species accept electrons from the electrode, which can be

observed as cathodic current. Oxidation occurs at anode, arising in anodic current; during

oxidation the electrode accepts electrons from the electroactive species. Intensity of the

faradaic currents corresponds to the quantity of the species and the potential at which they

occur provides information on the identity of the analyzed species. Species exhibiting surface

activity can undergo adsorption/desorption or reorientation processes which result in

tensammetric (nonfaradaic) currents (Bard and Faulkner, 2001).

3.1 Types of electrodes

In the three-electrode systems, working, reference and counter electrodes are used.

Reference electrodes exhibit known and stable potential (ideally where no charge transfer

24

over a wide potential range occurs) and are made of metal covered with a layer of its salt in a

solution containing an anion of this salt. Typical examples are the silver chloride electrode,

mercury-mercurous sulfate electrode or calomel electrode. Counter electrodes are made of

electrochemically inert materials such as platinum or carbon. Its surface area is often much

larger than that of the working electrode to ensure that the half-reaction occurring there can

proceed fast enough and does not limit the process at the working electrode.

Working electrodes are made of plenty of materials involving mercury, carbon, or

noble metals such as gold or platinum. Material of the working electrode strongly influences

its use and suitability for various electrochemical techniques. Working electrode should

provide high signal-to-noise characteristics, as well as a reproducible response. Thus, its

selection depends primarily on two factors: the redox behavior of the target analyte and the

background current over the potential region required for the measurement. Other

considerations include the potential window, electrical conductivity, surface reproducibility,

mechanical properties, cost and toxicity (Wang, 2001a). Potential window is usually limited

by electrolysis of water; anodic potential range is limited by the oxygen evolution, cathodic

potential range by the hydrogen evolution. Both these processes are pH-dependent.

3.1.1 Mercury electrodes

Mercury is a traditional material used in electroanalytical chemistry. Using mercury

electrodes is highly advantageous as their surface is atomically smooth, can be easily renewed

and is highly reproducible. Another advantage lies in the fact that mercury electrodes exhibit

high values of hydrogen overvoltage and therefore enable measurements at very negative

potentials. However, at mildly positive potentials, mercury is oxidized and the anodic range is

thus limited. Therefore, mercury electrodes are convenient for analysis of reduction processes.

Mercury electrodes are liquid and as such are not suitable for use in biosensors, where solid

materials are more favorable.

Various types of mercury electrodes are being used, the most common of which are

dropping mercury electrode (DME), hanging mercury drop electrode (HMDE) or mercury

film electrode (MFE). Besides these types of liquid mercury electrodes, solid amalgam

electrodes (SAE) have been used, which are alloys of mercury with another metal, such as

silver or copper. They constitute a non-toxic alternative of mercury electrodes, suitable for use

in sensors and exhibiting similar electrochemical characteristics as mercury electrodes such as

highly negative potentials enabling to observe reduction and catalytic hydrogen evolution.

25

SAE have been used in electrochemical analysis of many species, including DNA

(Yosypchuk et al., 2002; Fadrná et al., 2004; Hasoň and Vetterl, 2006).

3.1.2 Solid electrodes

Most frequently used solid working electrodes are made of various types of carbon or

a metal, such as gold or platinum. Each material has its own specific features but a common

characteristic of the solid electrodes (with some exceptions) is an anodic potential window

enabling analysis of oxidizable compounds. Compared to mercury electrodes, solid electrodes

exhibit good mechanical stability and therefore are suitable for use in portable devices. Solid

electrodes can be easily miniaturized and occur in many variants, such as rotating disk

electrodes or screen-printed electrodes. They offer inexpensive mass production and their

surface can be modified with a broad spectrum of films and layers.

An important factor in using solid electrodes is the dependence of the response on the

surface state of the electrode. Accordingly, the use of such electrodes requires precise

electrode pretreatment and polishing to obtain reproducible results. The nature of these

pretreatment steps depends on the materials involved. Mechanical polishing and potential

cycling are commonly used for metal electrodes, while various chemical or electrochemical

procedures are added for activating carbon-based electrodes. Unlike mercury electrodes, solid

electrodes present a heterogeneous surface with respect to electrochemical activity (Wang,

2001b).

Solid electrodes based on carbon are currently in widespread use in electroanalysis,

primarily because of their broad potential window, low background current, rich surface

chemistry, low cost, chemical inertness, and suitability for various sensing and detection

applications. Renewal of their surface is easier than in the case of metal electrodes. In

contrast, electron transfer rates observed at carbon surfaces are often slower than those

observed at metal electrodes. Microstructure of the electrode surface (edge- or basal-plane

orientation of the graphite sheets) has a profound effect on the electrochemical reactivity at

carbon electrodes as well as other factors, such as cleanliness of the surface and presence of

surface functional groups. The most commonly used carbon electrodes are:

- Pyrolytic graphite electrode (PGE) is made by heat treatment of pyrolytic carbon

or by chemical vapor deposition. PGE can occur in edge- or basal-plane

orientation. The basal-plane electrode consists of graphite layers which lie parallel

26

to the surface. In comparison, edge-plane electrodes are fabricated in a way that

the layers of graphite lie perpendicular to the surface (Banks and Compton, 2005).

- Glassy carbon electrode exhibits wide potential window, chemical inertness

(solvent resistance) and relatively reproducible performance. The structure of

glassy carbon involves thin, tangled ribbons of cross-linked graphite-like sheets.

The electrode material has a high density and small pore size and its surface can be

polished to achieve a shiny “mirror-like” appearance.

- Carbon paste electrode uses graphite powder mixed with various water-

immiscible nonconducting organic binders (pasting liquids such as mineral oil) and

offer an easily renewable and modified surface and very low background current

contributions. Disadvantage of carbon pastes is the tendency of the organic binder

to dissolve in solutions containing organic solvent.

- Diamond electrodes are fabricated by chemical vapor deposition methods. Since

diamond itself is an insulator, it has to be doped with boron to become conductive.

Diamond electrodes have wide potential window (approaching 4 V), low and

stable background currents and relatively low adsorption of organic molecules.

Platinum and gold electrodes are the most widely used metallic electrodes. Such

electrodes offer very favorable electron transfer kinetics and a large anodic potential range. In

contrast, the low hydrogen overvoltage at these electrodes limits the cathodic potential

window. Another limiting factor lies in the high background currents associated with the

formation of surface oxide or adsorbed hydrogen layers. Compared to platinum electrodes,

gold ones are more inert, and hence are less prone to the formation of stable oxide films or

surface contamination. Therefore the gold electrodes are often preferred in electroanalytical

chemistry. Gold electrodes are widely used as substrates for self-assembled organosulfur

monolayers, including thiol-derivatized DNA oligonucleotides (Herne and Tarlov, 1997).

3.2 Electrochemical methods

Since the discovery of polarography by Jaroslav Heyrovský in 1922, many

electrochemical methods have been presented, taking advantage of various approaches –

chronopotentiometry, chronoamperometry, voltammetry, or impedance spectroscopy. The

most frequently used method for DNA analysis is voltammetry. Voltammetry is based on the

same principle as polarography, the difference is in the fact that in polarography, the working

electrode is dropping mercury electrode (DME). During voltammetry (or polarography),

27

potential is applied to the working electrode and the actual current value is measured as the

dependent variable. The most common voltammetric methods are listed below.

3.2.1 Cyclic Voltammetry (CV)

Cyclic voltammetry is the most widely used voltammetric technique. The power of

cyclic voltammetry results from its ability to rapidly provide considerable information on the

thermodynamics of redox processes and the kinetics of heterogeneous electron transfer

reactions and on coupled chemical reactions or adsorption processes. Cyclic voltammetry is

often the first experiment performed in an electroanalytical study (Wang, 2001b).

Cyclic voltammetry consists of scanning linearly the potential of working electrode,

using a triangular potential waveform (Fig. 4). Depending on the information sought, single or

multiple cycles can be used. Scan rate (the rate of a potential change) is a very important

factor in CV. By adjusting a scan rate, nature of the electrode process and its reversibility can

be explored and it is possible to discriminate between redox, tensammetric or catalytic

processes.

Figure 4. Potential-time waveform used in cyclic voltammetry (three successive

potential cycles).

3.2.2 Linear Sweep Voltammetry (LSV)

Linear sweep voltammetry is very similar to cyclic voltammetry, the difference being

that in LSV a scan in only one direction is performed. The potential waveform thus resembles

that of CV but it lacks the reverse scan (Fig. 5). During the potential sweep, the potentiostat

measures the current resulting from the applied potential.

28

Figure 5. Potential-time waveform used in linear sweep voltammetry.

3.2.3 Differential-pulse voltammetry (DPV)

In differential-pulse voltammetry, fixed magnitude pulses (superimposed on a linear

potential ramp) are applied to the working electrode (Fig. 6). The current is sampled twice,

just before the pulse application and again late in the pulse life, when the charging current has

decayed. The first current is instrumentally subtracted from the second, and this current

difference is plotted against the applied potential. Introduction of pulse voltammetry methods

led to lowering detection limits as they increase the ratio between faradaic and non-faradaic

currents (Bond and Grabaric, 1979).

Figure 6. Potential-time waveform used in differential-pulse voltammetry.

3.2.4 Square-wave voltammetry (SWV)

Square-wave voltammetry is one of the pulse voltammetry methods. It is a differential

technique in which a waveform composed of a symmetric square wave, superimposed on a

base staircase potential, is applied to the working electrode (Fig. 7). The current is sampled

twice during each square-wave cycle, once at the end of the forward pulse and once at the end

of the reverse pulse. Since the square-wave modulation amplitude is very large, the reverse

pulses cause the reverse reaction of the product (of the forward pulse). The difference

between the two measurements is plotted versus the base staircase potential. Excellent

sensitivity results from the fact that the net current is larger than either the forward or reverse

29

components (since it is the difference between them); the sensitivity is higher than that of

differential pulse polarography (in which the reverse current is not used).

Figure 7. Potential-time waveform used in square-wave voltammetry.

3.2.5 Alternating current voltammetry (ACV)

Alternating current voltammetry is a frequency-domain technique which involves the

superimposition of a small amplitude AC voltage on a linear ramp (Fig. 8). Usually the

alternating potential has a frequency of tens to hundreds Hz and an amplitude of 10–20mV.

The AC signal thus causes a perturbation in the surface concentration, around the

concentration maintained by the DC potential ramp. The resulting AC current is plotted

against the potential. ACV is suitable for studies of adsorption processes.

Figure 8. Potential-time waveform used in alternating current voltammetry.

3.3 Electrochemistry of nucleic acids

Nucleic acids are electrochemically active, which was first discovered by Emil

Paleček in 1958 (Paleček, 1958, 1960) showing that it is possible to observe signals for

adenine, cytosine and guanine using oscillographic polarography at dropping mercury

electrode. First, only mercury electrode was used in DNA analysis but in 1970s it was shown

by Viktor Brabec and Glenn Dryhurst (Brabec and Dryhurst, 1978a; Brabec and Dryhurst,

1978b; Brabec, 1981) that DNA (bases adenine and guanine) yields analytically useful signals

at carbon electrodes, too. Another development in the field of nucleic acids electrochemistry

30

was achieved by introduction of adsorptive transfer stripping technique which led to reduced

volume of the sample required for analysis and increased sensitivity (Paleček and

Postbieglová, 1986).

3.3.1 Adsorptive transfer stripping in DNA analysis

Generally, stripping techniques are used for analyses of traces of metals. The

procedure is based on preconcentration of the analyte at the electrode surface and its

subsequent stripping. It is an extremely sensitive technique and it can lower the detection

limits by orders of magnitude compared to solution-phase voltammetric measurements

(Wang, 2001b). Besides metals, many inorganic and organic molecules that exhibit surface-

active properties can be analyzed using stripping techniques, including nucleic acids and

proteins. Depending on their redox activity, the adsorbed organic compounds can be observed

owing to their oxidation or reduction. Nonelectroactive macromolecules may also be

determined following their interfacial accumulation from tensammetric peaks.

In DNA analysis, adsorptive transfer stripping voltammetry (AdTSV) is often used,

which involves adsorption of DNA at the electrode surface from a small droplet of sample

(usually a few microliters) during open circuit, transfer of the electrode with the adsorbed

layer into a new medium containing blank electrolyte and subsequent voltammetric analysis

(Fojta et al., 2008). This approach takes advantage of the adsorption (strong enough to resist

exchange of the media) of DNA bases and sugar-phosphate backbone on surfaces of both

mercury and carbon electrodes (Brabec and Paleček, 1972; Brabec and Dryhurst, 1978a). This

technique has many advantages compared to conventional electrochemical analysis with the

analyte diluted in the background electrolyte. Probably the most important feature is that it

substantially reduces amount of the sample required for the analysis, which is often difficult

to obtain. Another advantage lies in the fact that DNA can be adsorbed from a solution which

differs from the electrolyte and is not suitable for electrochemical measurements and therefore

it is possible to separately optimize conditions for the adsorption and for the electrochemical

measurement. During the washing step possible interfering low molecular weight substances

can be removed. AdTSV also enables to exploit the differences in adsorbability of substances

on the electrode surface and to separate them accordingly. Moreover, it makes it possible to

study the interaction of biomacromolecules immobilized on the surface of the electrode with

substances contained in the solution without the results of the voltammetric measurement

being affected by the interactions in the bulk of the solution and to study the effect of

electrode potential on the properties and interactions of the adsorbed macromolecules

31

(Paleček and Postbieglová, 1986). Electrodes with adsorbed DNA layer can be used as simple

electrochemical biosensors (Paleček and Bartošík, 2012).

3.3.2 Analysis of DNA structure at mercury electrodes

When applying negative potentials at the mercury electrode, DNA adsorbed at the

electrode surface undergoes adsorption/desorption and reorientation processes which can be

observed as tensammetric peaks. Such peaks can provide valuable information on properties

of the studied DNA, especially its structure. For this purpose, alternating current voltammetry

(ACV) in weakly alkaline media is used.

At negative potentials, repulsion between negatively charged electrode surface and

negatively charged sugar-phosphate backbone of the DNA occurs, which can result in partial

desorption of the DNA molecules from the electrode surface. This leads in changes in

capacity of the electrode double layer, which can be observed as tensammetric signals. The

DNA structure influences which part of the DNA molecule participates in the

adsorption/desorption processes, which results in various tensammetric peaks (Fojta, 2004;

Palecek and Fojta, 2005).

At the AC voltammogram, three distinct tensammetric peaks can be observed. Peak 1

occurs at a potential around -1.2 V and is caused by desorption of the sugar-phosphate

backbone. At a potential around -1.4 V peak 3, which is caused by desorption of DNA bases,

can be observed. This peak occurs only if the DNA molecule is single stranded or contains

single stranded regions. Another tensammetric peak, peak 2, occurs at -1.3 V and corresponds

to changes in double stranded DNA during which distorted regions in the dsDNA molecule

are adsorbed via bases (Fojta et al., 1998; Fojta, 2004; Paleček and Bartošík, 2012).

The tensammetric signals can be applied as a sensitive tool for detection of DNA

damage. Covalently closed circular DNA such as supercoiled plasmid DNA (scDNA) does

not provide peak 3 since the bases are not accessible to the electrode surface. After

introduction of a strand break into DNA the DNA strands begin to unwind, which allows

adsorption of bases onto the electrode surface and peak 3 can thus be observed. This approach

enables to study various genotoxic agents and factors such as -radiation or hydroxyl radicals

and their influence on formation of single stand breaks (Fojta and Paleček, 1997).

3.3.3 DNA signals at mercury electrodes

Since their potential window is very wide at the negative side, mercury electrodes are

suitable for studying reduction of nucleic acids and their components. Adenine, cytosine, 5-

32

methylcytosine and guanine can be reduced at mercury electrodes. Adenine and cytosine are

reduced at very similar potential around -1.5 V (against Ag|AgCl|3M KCl reference electrode,

as all other potentials stated here), depending on pH of the electrolyte, and provide a common

cathodic peak CA (Paleček and Bartošík, 2012). Under certain conditions, especially when C

and A form homooligonucleotide blocks, it is possible to resolve their overlapping reduction

peaks using elimination voltammetry with linear scan (Trnková et al., 2003, 2006; Mikelová

et al., 2007). 5-methylcytosine is reduced at the same potential as cytosine, therefore it is not

possible to distinguish them when naturally occurring in DNA. Nevertheless, when cytosine

(unlike 5-methylcytosine) is converted to uracil using bisulfate treatment, decrease in C

reduction peak can be detected, which enables to observe levels of DNA methylation

(Bartošík et al., 2012).

Guanine can be reduced at highly negative potentials but its reduction peak cannot be

observed as it is overlapped by hydrogen background discharge. Guanine is reduced to 7,8-

dihydroguanine which can be oxidized during anodic scan in cyclic voltammetry providing

peak G at about -0.25 V (Trnková et al., 1980; Studničková et al., 1989). Recently it has been

proposed that electrochemically or electrocatalytically generated hydrogen radicals are

involved in chemical reduction of guanine to 7,8-dihydroguanine (Daňhel et al., 2016).

Thymine and uracil can be reduced at highly negative potentials at mercury electrodes

only in nonaqueous solution – it was shown that reduction of thymine (Cummings and Elving,

1979) and uracil (Cummings and Elving, 1978) can be observed in dimethylsulfoxide, which

is not convenient for the common DNA analysis.

Reduction signals of DNA bases at mercury electrodes are strongly dependent on the

DNA structure. While in single-stranded DNA (ssDNA) cytosine and adenine are accessible

to the electrode surface, in double-stranded DNA (dsDNA) their reduction sites are hidden

inside the double helix as they participate in the Watson-Crick base pairing (Fig. 9). This

results in substantially lower CA peak in dsDNA compared to ssDNA. The same applies to

guanines when forming G-quadruplexes. Hoogsteen-paired guanine residues in the

quadruplexes have only limited accessibility for the reduction process at the negatively

charged surface of the mercury electrode, which can be observed as decrease in the G peak

intensity when G-quadruplexes are formed (Vidláková et al., 2015).

33

Figure 9. Location of primary reduction sites of G, C and A at mercury electrodes (magenta

rectangles) and primary oxidation sites of G and A at carbon electrodes (blue circles).

3.3.4 DNA signals at carbon electrodes

Carbon electrodes are (unlike mercury electrodes) suitable for studying oxidation

processes when applying positive potentials. Adenine and guanine are oxidized at carbon

electrodes and they provide oxidation signals at potentials about 1.1 V for guanine and 1.3-1.4

V for adenine, which was first discovered in the 1970s (Dryhurst and Pace, 1970; Brabec and

Dryhurst, 1978a, 1978b; Brabec, 1981). These signals are the most frequently analytically

used signals provided by DNA at carbon electrodes (Paleček and Bartošík, 2012). Pyrimidine

bases can be also oxidized but their oxidation occurs at more positive potentials and the

signals thus interfere with the background currents caused by anodic oxygen evolution, which

hinder the detection of the oxidation peaks (Brotons et al., 2016). However, several

approaches have been developed that enable simultaneous detection of oxidation of all four

natural nucleotides, involving use of differential pulse voltammetry at glassy carbon electrode

(Oliveira-Brett et al., 2004) or use of chemically reduced graphene oxide modified glassy

carbon electrode, which enables to study DNA bases in both single-stranded DNA and

double-stranded DNA without a prehydrolysis step (Zhou et al., 2009).

Recently it has been shown in our lab for the first time that carbon electrodes, namely

pyrolytic graphite electrode in basal plane orientation, are suitable for studying not only

anodic oxidation processes but also cathodic reduction of the nucleobases up to potential of -

2.0 V. Moreover, products of irreversible oxidation/reduction of the parent bases were shown

34

to yield analytically useful, base-specific cathodic/anodic signals, making it possible to

distinguish between the canonical bases (adenine, cytosine, guanine and thymine), uracil and

5-methylcytosine in DNA without need of DNA hydrolysis or electrode modification, which

makes it an excellent tool for simultaneous label-free analysis of bases in DNA (Špaček et al.,

2017).

Although they are not natural DNA bases, it is worth mentioning that 7-deazapurines

(analogs of natural purine bases in which N7 atom is replaced by CH group) provide

oxidation signals at carbon electrodes, too. Both 7-deazaguanine (G*) and 7-deazaadenine

(A*) exhibit significantly lower potentials of their oxidation, compared to the respective

natural nucleobases. G* is oxidized at a potential around 0.8 V, which is about 300 mV less

positive than oxidation signal of G, A* at a potential around 1.1 V, which is also about 200-

300 mV less positive than potential at which A is oxidized and moreover, it overlaps with the

oxidation peak of G (Pivoňková et al., 2010). At mercury electrodes, A* can be reduced

giving rise to a similar irreversible cathodic peak as the natural A and G* does not yield any

peak analogous to the peak G due to guanine, in agreement with a loss of corresponding redox

site in G* (Dudová et al., 2016).

Generally, oxidation peaks gained at carbon electrodes are not as sensitive to changes

in DNA structure as reduction and tensammetric peaks obtained at mercury electrodes

(Paleček and Bartošík, 2012; Brotons et al., 2016).

3.4 Electrochemical labeling of nucleic acids

Although DNA possesses intrinsic electroactivity and their redox and tensammetric

signals can be used for various analytical applications, sometimes it is advantageous to use

DNA labeling by various chemical moieties which can be electrochemically detected. Using

electroactive DNA labeling has the following advantages compared to analysis of natural

DNA:

- Using DNA labels increases both sensitivity and selectivity of the analysis (Fojta

et al., 2007; Hocek and Fojta, 2011), enabling detection of the labeled DNA

species in the sample even if the unlabelled DNA is overabundant.

- DNA labels usually provide their signals at potentials that are not as extreme as in

the case of DNA bases, which are reduced or oxidized at very high negative or

positive potentials (Paleček and Bartošík, 2012).

35

- Unlike unlabeled DNA, which has been analyzed practically only at mercury- and

carbon-based electrodes, the DNA labels offer a wider range of possibilities as to

which electrode material can be used for the analysis. For example gold electrodes,

not very suitable for measuring DNA as only guanine signal can be studied

(Ferapontova and Domínguez, 2003), can be used for construction of DNA

hybridization sensors using probe oligonucleotides immobilized at the electrode

surface using thiol linkers (Flechsig and Reske, 2007; Surkus and Flechsig, 2009;

Jacobsen and Flechsig, 2013).

DNA can be labeled in various ways: DNA labels can either noncovalently interact

with DNA through groove binding or intercalation, or can covalently bind accessible reactive

groups in DNA (or RNA) such as in the case of osmium tetroxide complexes (Flechsig and

Reske, 2007; Fojta et al., 2007, 2011; Trefulka et al., 2007; Havran et al., 2008). Another

possibility lies in enzymatic incorporation of electrochemically modified nucleotides (Hocek

and Fojta, 2011). Many applications have been presented that utilize DNA labeling, with the

most common ones being used in construction of DNA hybridization sensors (Jelen et al.,

2002; Fojta et al., 2003, 2004; Fojta et al., 2007; Horáková et al., 2011), sensors of DNA

structure (Palecek and Hung, 1983; Paleček, 1992a) or DNA damage (Fojta, 2002; Fojta et

al., 2016) and in detection of DNA-proteins interactions (for more details see section 2.4).

3.4.1 Noncovalently bound redox indicators

Generally, noncovalently bound indicators are structure-specific and therefore suitable

for discrimination of ssDNA and dsDNA immobilized at the electrode surface. They can bind

DNA in various modes: indicators with cationic nature that interact electrostatically with the

polyanionic DNA chain; indicators binding to DNA groove, such as Hoechst 33258; planar

aromatic molecules capable of intercalation between adjacent bases in the DNA doublehelix,

such as echinomycin (Jelen et al., 2002), actinomycin D and proflavine (Gebala et al., 2009)

or anthraquinone (Wong and Gooding, 2003); bisintercalators and threading intercalators with

increased selectivity for dsDNA compared to regular intercalators.

3.4.2 Osmium tetroxide complexes

Complexes of osmium tetroxide with nitrogenous ligands (Os,L) were the first used

covalently binding electroactive DNA labels. Osmium tetroxide reacts with C=C double

bonds. The process involves [3+2] addition of osmium tetroxide across the C=C double bond

giving rise to an osmic acid diester (glycolate), that is subsequently hydrolyzed to the glycol

moiety and osmate. Analogous reactions are given by various compounds possessing the C=C

36

double bonds, including pyrimidine nucleobases (at C5=C6) and indole moiety featuring side

group of amino acid tryptophan (W, at C2=C3). It has been established that tertiary amines

stabilize the osmium(VI) glycolates upon coordination of the central osmium atom by the

nitrogenous ligands. Thus, products of modification of the pyrimidine or W residues with

osmium tetroxide complexes bearing the nitrogenous ligands are stable adducts retaining the

osmium moiety and the given ligand (Paleček, 1992b; Deubel, 2003; Fojta et al., 2011).

Pyridine (py) was the first nitrogenous ligand used in complexes with osmium tetroxide, later

more ligands have been presented such as 2,2’-bipyridine (bpy), 1,10- phenanthroline (phen)

derivatives or N,N,N’,N’-tetramethyl ethylenediamine (TEMED). These ligands influence the

potentials at which the Os,L complexes provide voltammetric signals, which was used in

“multicolor” DNA labeling enabling parallel analysis of multiple DNA targets (Fojta et al.,

2007).

In DNA, osmium tetroxide complexes react preferentially with pyrimidine

nucleobases, with thymine being most reactive, followed by uracil and cytosine (Reske et al.,

2009). The reaction proceeds only when the C5=C6 double bond is accessible to the Os,L

complex, which happens only in the case of single-stranded DNA. This makes Os,L

complexes excellent tool for probing DNA structure as they can very precisely discriminate

between ssDNA and dsDNA (Jelen et al., 1991; Paleček, 1992b). Six-valent osmium in

complex with nitrogenous ligands has also been used for labeling nucleic acids but instead of

reacting with pyrimidine nucleobases, it condensates with cis-diols of sugar residues, which

narrows it down (in the sense of NA labeling) to labeling of RNA. This was utilized for

labeling ribose at the 3’ end of RNA oligonucleotides (Trefulka et al., 2010).

Adducts of Os,L with DNA can be detected using different types of electrodes due to

the electrochemical activity of the central osmium atom, which can undergo several redox

processes. At mercury-based electrodes, Os,bpy modified DNA provides three reversible

faradaic peaks between 0 and -1 V and a catalytic peak, which appears close to the hydrogen

background discharge and is caused by catalytic hydrogen evolution by Os,L (Fig. 10). This

peak allows determination of labeled DNA in very low concentrations (Palecek and Hung,

1983; Yosypchuk et al., 2006; Paleček et al., 2009; Fojta et al., 2011). At carbon electrodes,

osmium tetroxide complexes also yield faradaic signals, which are analytically useful (Fojta et

al., 2003; Fojta et al., 2007). There is a sufficient difference between the potentials of free

Os,bpy and Os,bpy-DNA adducts, which enables to determine labeled DNA even in the

presence of unreacted Os,bpy when using an adsorptive transfer stripping voltammetric

37

procedure involving extraction of free Os,bpy from the electrode by chloroform (Fojta et al.,

2002).

Figure 10. Schematic representation of redox potentials of natural DNA bases, 7-

deazapurines and examples of electroactive moieties coupled to the nucleobases (Hocek and

Fojta, 2011).

3.4.3 Redox labels and their enzymatic incorporation

Another option for introduction of covalently bound label lies in enzymatic

incorporation of base-modified nucleotides into DNA. It has been shown already in 1981 that

DNA and RNA polymerases are able to incorporate modified nucleotides into the growing

chain (Langer et al., 1981). There are many polymerases, which have been tested for

incorporation of diverse modified nucleoside triphosphates (dNXTPs). In general, B-family

thermo-stable polymerases (Vent (exo-), Pwo, KOD) are the best enzymes tolerating the

38

presence of most modifications, A-family polymerases are much less suitable (Hocek and

Fojta, 2011). Apart from that, other enzymes have been found applicable for incorporation of

modified dNTPs, such as terminal deoxynucleotidyl transferase (TdT, for more details see

section 2.3) or reverse transcriptase (Adelfinskaya and Herdewijn, 2007).

Several methods can be used for incorporation of dNXTPs into DNA, such as primer

extension (PEX) during which the DNA polymerase incorporates nucleotides at the 3’-OH

end of the growing primer in the presence of a longer template. When one (or more) of the

natural dNTPs is replaced by a modified one (dNXTP), the polymerase incorporates the

modification at the position of the particular nucleobase. Using this method, one or several

modifications can be introduced at the end of the primer strand (Riedl et al., 2009; Hocek and

Fojta, 2011; Balintová et al., 2013; Simonova et al., 2014). Other methods suitable for

incorporation of dNXTPs into DNA comprise polymerase chain reaction (PCR) (Jäger et al.,

2005; Čapek et al., 2007), Nicking Enzyme Amplification Reaction (NEAR) (Menova et al.,

2013) or TdT tailing reaction (Arzumanov et al., 2000; Anne et al., 2007; Horáková et al.,

2011; Havranová-Vidláková et al., 2018).

Synthesis of the dNXTPs is done either in a classical way consisting of the synthesis of

a modified nucleoside followed by chemical triphosphorylation or by modification of

halogenated dNTPs by aqueous-phase cross-coupling reactions, which is often more efficient.

Modifications are usually linked to C5 of pyrimidine nucleotides or to C7 of 7-deazapurine

nucleotides because modifications in these positions affect DNA structure and function less

than in other positions. For example 8-substituted purine dNTPs are poor substrates for

polymerases. Modifications are either attached directly or via acetylene or phenylene linkers

(Hocek and Fojta, 2008, 2011).

The redox labels used for modification of DNA are chosen with respect to their

electroactivity, which means that their signals should not interfere with reduction or oxidation

signals of natural nucleobases or with the background electrolyte discharge. The redox

processes often include multiple electrons and can be reversible. Examples of suitable redox

labels (for some of them, see Fig. 10) used for enzymatic incorporation into DNA in our lab

in cooperation with Michal Hocek’s lab at IOCB Prague, are: ferrocene (Brázdilová et al.,

2007), Os(bpy)3 and Ru(bpy)3 complexes (Vrábel et al., 2009), aminophenyl and nitrophenyl

(Cahová et al., 2008), tetrathiafulvalene (Riedl et al., 2009), anthraquinone (Balintová et al.,

2011), benzofurazane (Balintová et al., 2013), methoxyphenol and dihydrobenzofuran

39

(Simonova et al., 2014), azidophenyl (Daňhel et al., 2016) and phenothiazine (Simonova et

al., 2017).

40

3. Aims of the dissertation

1) Development of a new method for detection of protein-DNA interactions based on dual

redox labeling of DNA.

2) Electrochemical study of products of terminal deoxynucleotidyl transferase tailing

reaction.

3) Modification of butylacrylate conjugates in single- and double-stranded DNA with

osmium tetroxide complex.

41

4. Results and discussion

4.2 Label-free voltammetric detection of products of terminal

deoxynucleotidyl transferase tailing reaction

Monika Hermanová, Pavlína Havranová-Vidláková, Anna Ondráčková, Swathi Senthil

Kumar, Richard Bowater, Miroslav Fojta

Appendix 2

In this work, terminal deoxynucleotidyl transferase (TdT) tailing reaction has been studied

using electrochemical methods. For the TdT tailing reaction, three types of

homooligonucleotides – A30, C30 and T30 – were used as primers to which various dNTPs

(dATP, dCTP, dGTP, dTTP, dA*TP and dG*TP or their combinations) were added to form

the tail. The tailing reaction was performed in five different primer:dNTP molar ratios: 1:1,

1:5, 1:10, 1:20 and 1:50; no enzyme control was performed in order to compare signals

gained for the tailing reaction products to those of the sole primer.

Guanine (G) and 7-deazaguanine (G*) were used for the tailing reaction with all three primers

- A30, C30 and T30 (Figure 16). A30+G combination revealed a significant difference between

the results gained at HMDE and PGE for the same reaction products (Fig. 16A). At PGE,

height of the GOX

peak increased with the increasing primer:dNTP ratio. In contrast, signals

gained at the HMDE showed a decreasing tendency from 1:5 to 1:50 ratio after the initial

growth between 1:1 and 1:5 ratio. This behavior at HMDE in contrast to what was observed at

PGE has been previously observed and was ascribed to formation of G-quadruplexes (G4)

(Vidláková et al., 2015). The denaturing PAGE autoradiogram also indicated formation of

G4. The formation of G4 structures has two aspects. Firstly, when the primer:nucleotide ratio

is sufficient to produce G-tails capable of G4 formation, the tailing reaction is stopped when

the G4 is formed because of the inaccessibility of the 3’-OH end to TdT. Secondly, the G4

formation changes behavior of the oligonucleotides at the mercury electrode surface since

mercury-containing electrodes are sensitive to changes in DNA structure. On the other hand,

oxidation of guanine at PGE depends merely on the length of the oligonucleotide molecules

and does not reflect any structural changes occurring in the molecules. Electrochemical

analysis at PGE thus provided comparable results as PAGE analysis. In the case of T30+G

(Figure 16C), the PGE and denaturing PAGE revealed the same pattern of the length of the

42

tailing reaction products as A30+G, indicating formation of G4, however these results could

not be complemented with analysis at HMDE because of prevailing behavior of T30

oligonucleotide at the mercury electrode surface. Unlike A30+G and T30+G, the C30+G tailing

reaction products (Figure 16B) did not form G4, and formed another DNA structure – DNA

hairpin since bases C and G are complementary. For comparison with behavior of the natural

guanine, 7-deazaguanine (G*) was used, which is an analogue of natural guanine incapable of

forming multi-stranded DNA structures such as triplexes or quadruplexes. Both voltammetric

analysis at PGE and PAGE analysis (Figure 16D) showed that, unlike in the case of A30+G

tailing products, length of the tailing products grew gradually with the rising ratio, which was

caused by the fact that G* cannot form G4. Practically the same results as for the A30+G*

were obtained for the T30+G* tailing products (Figure 16F) with similar pattern of the tailing

products lengths visible at PAGE autoradiogram and intensities of the G*OX

signal gained at

PGE. On the other hand, the pattern observed for C30+G* (Figure 16E) resembled more that

of the C30+G tailing products than those of A30+G* and T30+G*, which is in accordance with

the fact that the ability of G* to pair with C is not violated, which leads to hairpin formation

after a certain length of the tail is reached.

43

Figure 16. Peak heights of the respective peaks gained at HMDE or PGE and denaturing

PAGE autoradiograms obtained for combinations of G and G* with all three primers: (A)

A30+G; (B) A30+C; (C) T30+G; (D) A30+G*; (E) C30+G*; (F) T30+G*.

Besides G and G*, adenine and 7-deazaadenine (A*) were also used for the tailing reactions

together with primers T30, C30 and A30. For the T30+A (Figure 17A), PAGE analysis revealed

that all reaction products for the ratios 1:5, 1:10, 1:20 and 1:50 had tails consisting of 5 or 6

nucleotides. It is expected that this behavior was caused by a formation of a DNA hairpin

since A and T are complementary bases. When the A tail reaches sufficient length for the

hairpin formation, the DNA bends and a DNA hairpin with a 5' overhang is created. This

means that the 3'-OH end becomes inaccessible for the TdT and the enzyme cannot continue

in the tailing reaction. Electrochemical results obtained for the T30+A again displayed

significant difference between the data obtained at HMDE and PGE (similarly to A30+G and

C30+G). The C30+A combination (Figure 17B) is specific as to the signal gained for the bases

forming the primer and the tail because in DNA with random distribution of bases, C and A

44

yield a common CA peak at HMDE since the potentials at which they are reduced are in close

proximity. However, in DNA composed of C and A homooligonucleotide stretches, two

separated peaks can be observed under certain conditions (Trnková et al., 2003, 2006;

Mikelová et al., 2007). In this case, it was not possible to see the separated C and AHMDE

peaks, only a hint of the AHMDE

peak appeared at a potential more negative than that of the C

peak for the product with the longest A tail (the 1:50 ratio). At PGE, the intensity of the AOX

peak grew with the growing ratio but it was not possible to compare the trend of this signal to

the pattern gained at PAGE since in this case, the PAGE analysis did not enable to determine

lengths of the tailing products. Most probably it was caused by a wider distribution of the

lengths leading to smaller amount of DNA of one specific length meaning that the band

corresponding to such length could not be visualized. Here, we can thus see that the

electrochemical analysis at PGE can determine length of the tailing products more precisely

than the PAGE analysis. The same effect can be observed at the PAGE autoradiogram of the

C30+A* tailing products (Figure 17C), again making the electrochemical analysis at PGE

more accurate than the PAGE analysis. Results gained for the A30+A* tailing products (Figure

17D) show that the trend of the growing tail length visible at both PAGE autoradiogram and

electrochemical analysis at PGE is different from that of C30+A* tailing products but again

these two methods provide comparable results.

45

Figure 17. Peak heights of the respective peaks gained at HMDE or PGE and denaturing

PAGE autoradiograms obtained for combinations of A and A* with all three primers: (A)

T30+A; (B) C30+A; (C) C30+A*; (D) A30+A*.

PAGE analysis of tailing products having tails consisting of cytosines (T30+C – Fig. 18A and

A30+C – 18B) showed that the length of the products grows with the growing ratio but even at

the highest ratios, the products are not as long as for other bases. Cyclic voltammetry at

HMDE was used for their analysis. For T30+C, a tensammetric peak caused by the T30 primer

(TTENS

) was observed but although it appeared at a potential close to that of the C peak, it was

possible to distinguish these two peaks as the potential at which it occurred was about 400

mV more negative than that of the C peak. In the case of A30+C, similar situation occurred as

for the C30+A, with C and A yielding a common reduction CA peak at HMDE. However, here

it was possible to determine the C peak for the higher ratios. With a growing number of

cytosines forming the tail, it was observed that the C peak appeared at a less negative

potential than the AHMDE

peak. For the 1:50 ratio, the peak was already well developed. The

peak heights gained for the CA peak showed no obvious trend and remained similar for the

primer and all the ratios. Nevertheless, an evident tendency to grow with the increasing ratio

was seen when the C peak was analyzed separately. Primers A30 and C30 were also used for

tailing reaction with thymines. The PAGE analysis of the A30+T tailing products (Fig. 18C)

showed that unlike for T30+A, the tailing reaction was not stopped after a few added

nucleotides and the tailing products gradually grew up to the 1:50 ratio. When analyzed at

HMDE, height of the A peak displayed decreasing tendency, which reflects decreasing

amount of the A30 parts of the tailing products that were adsorbed at the electrode surface due

to growing portion of the T-tails. The PAGE analysis of the C30+T (Fig. 18D) revealed the

same problem as for the C30+A and C30+A* combinations – no visible tailing products at the

autoradiogram. The voltammetric analysis of C30+T at HMDE confirmed that the cytosines

adsorb very strongly at the mercury surface (Hasoň et al., 2008), which is evident from the

fact that the C peak height remains unchanged for all the ratios.

46

Figure 18. Peak heights of the respective peaks gained at HMDE and denaturing PAGE

autoradiograms obtained for combinations of C and T with all three primers: (A) T30+C; (B)

A30+C; (C) A30+T; (D) C30+T.

47

4.3 Butylacrylate-nucleobase Conjugates as Targets for Two-step

Redox Labeling of DNA with an Osmium Tetroxide Complex

Pavlína Havranová-Vidláková, Jan Špaček, Lada Vítová, Monika Hermanová, Jitka

Dadová, Veronika Raindlová, Michal Hocek, Miroslav Fojta and Luděk Havran

Appendix 3

In this work, butylacrylate- (BA) nucleobase (7-deaza adenine or uracil) conjugates

incorporated into DNA oligonucleotides have been tested for use in modification with

osmium tetroxide complex (Os,bpy). Single-stranded oligonucleotides bearing the dNBA

conjugates were prepared using terminal deoxynucleotidyl transferase (TdT) tailing reaction

(Figure 19, path a). In contrast to primer extension (PEX) technique with a template

dependent DNA polymerase (Figure 19, path b), which was successfully applied to construct

the dNBA

-modified DNA in our previous work (Dadová et al., 2013), the TdT tailing

technique has been tested here for the first time.

Fig. 19. Scheme of introduction of a single dNBA

nucleotide at the 3’-terminus of a

single-stranded initiator (path a) or templated PEX synthesis of ds ON bearing four dNBA

nucleotides (path b); and osmylation of the acrylate moieties in ss and ds DNA.

48

Denaturing polyacrylamide gel electrophoresis analysis showed that TdT tailing

reaction was successful in adding both unlabeled and BA-labeled dNTPs to the dA20 primer

(Figure 20). Ex situ AdTS SWV at PGE was used to analyze electrochemically the TdT

tailing products after treating them with 2 mM Os,bpy (Figure 20b,c). The dA20 initiator, in

agreement with the absence of any Os,bpy-reactive nucleobases in it, did not give any adduct-

specific peak and the same applies to the TdT tailing product with dATP (Fig, 20b). In

contrast, all the other TdT products (with dT, dU, dA*, dABA

and dUBA

) yielded well

developed signals in the region of peak , which differed in peak potentials as well as in

intensities. Peak heights obtained after Os,bpy modification of individual TdT tailing

products exhibited significant differences, with the most intense response recorded for dABA

and only negligible peak a observed for dU (Figure 20b,c). For both dABA

and dUBA

conjugates, significant increase of the signal was detected when compared to the respective

parent nucleotides dA* or dU, suggesting a considerable reactivity for the acrylate moiety

and/or modification of both C7=C8 and C2’’=C3’’ double bonds in the dABA

conjugate.

49

Fig. 20. a) Denaturating PAGE analysis of TdT tailing products of dA20 initiator and

individual dNTPs (molecular ratio 1:2) as indicated in the Figure; “A20” stands for the

unextended initiator. b) Comparison of voltammetric responses of Os,bpy-treated TdT

elongation products at PGE. c) Comparison of peak a heights for the TdT elongation products.

As the evaluation of these data was limited because of varying length of the TdT

tailing reaction products, ds ONs bearing dABA

by means of the PEX reaction were prepared

for the following experiments. Each molecule of the PEX products contained four dABA

conjugates or the parent dA* nucleotides. Figure 21 shows a dramatic difference between ds

PEX products bearing dABA

and those involving dA*. Peak heights continuously increased

with Os,bpy concentration between 0.5 and 5 mM (during 30-min reaction at 10°C) for the

dABA

PEX product, while no modification specific signal was detected for dA* up to 3 mM

Os,bpy under the same conditions. Only after treatment with 5 mM Os,bpy the dA*-

containing dsDNA exhibited some degree of modification (Figure 21a). When the dsDNA

with incorporated dABA

was incubated with 2 mM Os,bpy, peak height increased sharply up

50

to 20– 30 min and then levelled off (Figure 21b). Under the same conditions, no sign of the

dA* control DNA modification was detected at least up to 60-min incubation. Thus, under

milder conditions (such as 2 mM Os,bpy, 30-min modification) nucleobases in dA*-

containing dsDNA (similarly as those in fully natural dsDNA (Paleček, 1992b)) were

protected from Os,bpy modification, while the BA moiety in dABA

clearly exhibited

considerable reactivity even in the dsDNA.

Fig. 21. Os,bpy modification of double-stranded PEX products with incorporated dA*

or dABA

. Dependence of peak height a) on concentration of the Os,bpy reagent; time of

modifition 30 min; b) on time of modification; 2 mM Os,bpy. T=10°C.

51

5. Conclusions

We have presented three novel applications based on electrochemistry of DNA, taking

advantages of either intrinsic signals provided by DNA at mercury and carbon electrodes or

labeling of DNA with various electroactive markers.

First application can detect interactions of p53 protein with DNA. The p53 protein was

used as model protein as it can bind DNA in various modes therefore its use is appropriate for

development of a new methodology for studying DNA-protein interactions. Moreover its

DNA binding modes are very well described and especially in our laboratory, its DNA

binding properties have been intensively studied. The developed approach is based on dual

labeling of DNA probes and magnetic beads-based immunoprecipitation assay with

subsequent electrochemical detection. The DNA probes were labeled with two redox markers

– one label was used for a DNA probe bearing a specific binding site for the p53 protein

(p53CON), second label for a DNA probe lacking such site. We showed that using the

competition arrangement with these two labeled DNA probes, the specific and non-specific

binding of the p53 protein could be easily distinguished. The competition binding experiment

can thus provide information on relative affinities of the protein to the different DNA probes.

Furthermore, the p53 binding to DNA can be modulated by specific monoclonal antibodies.

This approach can be, after selection of appropriate DNA probes and suitable monoclonal

antibodies, potentially applied to study DNA-binding properties of other proteins.

Second approach comprised monitoring of terminal deoxynucleotidyl transferase

tailing reactions using intrinsic electroactivity of nucleobases. Various combinations of

homooligonucleotide primers and dNTPs were used for the tailing reactions and signals of the

respective bases forming the tail were studied using hanging mercury drop electrode (HMDE)

and pyrolytic graphite electrode (PGE). PGE showed to be very suitable for precise

determination of the tailing reaction products length as it corresponded to results obtained

using denaturing polyacrylamide gel electrophoresis (PAGE) and in some cases it even

outperformed the PAGE analysis. However, this approach enabled not only to study the

tailing reaction as such but also to observe formation of DNA structures, which was possible

due to use of HMDE for the analysis since it is very sensitive to DNA structure. Signals

obtained for some of the tailing products at HMDE showed decreasing trend, which reflected

structural changes occurring in the tailing reaction products and formation of various DNA

structures, such as DNA hairpins and G-quadruplexes. Thus, this approach proves to be an

52

excellent tool for studying the TdT tailing reactions but also for exploring electrochemical

behavior of the DNA oligonucleotides and influence of their structure at electrode surfaces.

In the third application, a new two-step technique of DNA modification with the

electroactive osmium tetroxide complex (Os, bpy) has been developed. The technique consists

of polymerase construction of DNA bearing butylacrylate (BA) moieties attached to uracil or

to 7-deaza adenine, followed by chemical modification of a reactive C=C double bond in the

acrylate residue. The BA-functionalized oligonucleotides were prepared either by TdT tailing

reaction or primer extension reaction. This approach enabled modification of the BA

conjugates in both single- and double-stranded (ds) oligonucleotides under conditions when

modification within the nucleobase rings in ds DNA was not possible. Moreover, Os,bpy

adducts formed with particular nucleobases and the BA conjugates can be electrochemically

distinguished on the basis of different redox potentials. These results are promising for the

application of the general two-step procedure in redox labeling of double-stranded DNAs and

in studies of DNA and interactions changing the accessibility of reactive groups located

within the DNA major groove.

53

6. References

Adelfinskaya, O. and Herdewijn, P. (2007) ‘Amino acid phosphoramidate nucleotides as

alternative substrates for HIV-1 reverse transcriptase’, Angewandte Chemie - International

Edition, 46(23), pp. 4356–4358. doi: 10.1002/anie.200605016.

Anne, A. et al. (2007) ‘Enzymatic redox 3’-end-labeling of DNA oligonucleotide monolayers

on gold surfaces using terminal deoxynucleotidyl transferase (TdT)-mediated single base

extension’, J Am Chem Soc, 129(10), pp. 2734–2735. doi: 10.1021/ja067954v.

Arzumanov, A. A. et al. (2000) ‘Terminal deoxynucleotidyl transferase catalyzes the reaction

of DNA phosphorylation.’, Nucleic acids research, 28(5), pp. 1276–81. Available at:

http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=102600&tool=pmcentrez&rendert

ype=abstract%5Cnhttp://www.ncbi.nlm.nih.gov/pubmed/10666473%5Cnhttp://www.pubmed

central.nih.gov/articlerender.fcgi?artid=PMC102600.

Bai, L. and Zhu, W. (2006) ‘p53 : Structure , Function and Therapeutic Applications’, Journal

of Cancer Molecules, 2(4), pp. 141–153. doi: 10.1007/978-94-017-9211-0_16.

Bakalkin, G. et al. (1995) ‘p53 binds single-stranded DNA ends through the C-terminal

domain and internal DNA segments via the middle domain’, Nucleic Acids Research, 23(3),

pp. 362–369. doi: 10.1093/nar/23.3.362.

Balintová, J. et al. (2011) ‘Anthraquinone as a redox label for DNA: Synthesis, enzymatic

incorporation, and electrochemistry of anthraquinone-modified nucleosides, nucleotides, and

DNA’, Chemistry - A European Journal, 17(50), pp. 14063–14073. doi:

10.1002/chem.201101883.

Balintová, J. et al. (2013) ‘Benzofurazane as a new redox label for electrochemical detection

of DNA: Towards multipotential redox coding of DNA bases’, Chemistry - A European

Journal, 19(38), pp. 12720–12731. doi: 10.1002/chem.201301868.

Balintová, J. et al. (2015) ‘Azidophenyl as a click-transformable redox label of DNA suitable

for electrochemical detection of DNA–protein interactions’, Chem. Sci., 6(1), pp. 575–587.

doi: 10.1039/c4sc01906g.

Baltimore, D. (1974) ‘Is terminal deoxynucleotidyl transferase a somatic mutagen in

lymphocytes?’, Nature, 248, pp. 409–411.

Ban, C. et al. (2004) ‘Detection of protein-DNA interaction with a DNA probe: distinction

between single-strand and double-strand DNA-protein interaction.’, Nucleic acids research,

32(13), p. e110. doi: 10.1093/nar/gnh109.

Banks, C. E. and Compton, R. G. (2005) ‘Edge Plane Pyrolytic Graphite Electrodes in

Electroanalysis: An Overview’, Analytical Sciences, 21(11), pp. 1263–1268. doi:

10.2116/analsci.21.1263.

Bard, A. J. and Faulkner, L. R. (2001) Electrochemical methods: fundamentals and

applications, JohnWilley & Sons.

Bartošík, M., Fojta, M. and Paleček, E. (2012) ‘Electrochemical detection of 5-

54

methylcytosine in bisulfite-treated DNA’, Electrochimica Acta, 78, pp. 75–81. doi:

10.1016/j.electacta.2012.05.115.

Basu, M., Hegde, M. V. and Modak, M. J. (1983) ‘Synthesis of compositionally unique DNA

by terminal deoxynucleotidyl transferase’, Biochemical and Biophysical Research

Communications, 111(3), pp. 1105–1112. doi: 10.1016/0006-291X(83)91413-4.

Bates, S. and Vousden, K. H. (1996) ‘P53 in Signaling Checkpoint Arrest or Apoptosis’,

Current Opinion in Genetics and Development, 6(1), pp. 12–18. doi: 10.1016/S0959-

437X(96)90004-0.

Bollum, F. J. (1960) ‘Calf thymus polymerase.’, The Journal of biological chemistry, 235(8),

pp. 2399–2403.

Bollum, F. J. (1974) ‘Terminal Deoxynucleotidyl Transferase’, Enzymes, 10, pp. 145–171.

doi: 10.1016/S1874-6047(08)60137-7.

Bond, A. M. and Grabaric, B. S. (1979) ‘Fast Sweep Differential Pulse Voltammetry at a

Dropping Mercury Electrode with Computerized Instrumentation’, Analytical Chemistry,

51(1), pp. 126–128.

Boulé, J. B., Rougeon, F. and Papanicolaou, C. (2001) ‘Terminal Deoxynucleotidyl

Transferase Indiscriminately Incorporates Ribonucleotides and Deoxyribonucleotides’,

Journal of Biological Chemistry, 276(33), pp. 31388–31393. doi: 10.1074/jbc.M105272200.

Bourdon, J. C. et al. (2002) ‘Scotin, a novel p53-inducible proapoptotic protein located in the

ER and the nuclear membrane’, Journal of Cell Biology, 158(2), pp. 235–246. doi:

10.1083/jcb.200203006.

Bowater, R. P. et al. (2015) ‘Biophysical and electrochemical studies of protein–nucleic acid

interactions’, Monatshefte für Chemie - Chemical Monthly, 146(5), pp. 723–739. doi:

10.1007/s00706-014-1405-4.

Brabec, V. (1981) ‘Nucleic Acid Analysis by Voltammetry at Carbon Electrodes’,

Bioelectrochemistry and Bioenergetics, 128(8), pp. 437–449. doi: 10.1016/0302-

4598(81)80005-0.

Brabec, V. and Dryhurst, G. (1978a) ‘Electrochemical Behavior of Natural and Biosynthetic

Polynucleotides at Pyrolytic-Graphite Electrode a New Probe for Studies of Polynucleotide

Structure and Reactions’, Journal of Electroanalytical Chemistry, 89(1), pp. 161–173. doi:

Doi 10.1016/S0022-0728(78)80041-2.

Brabec, V. and Dryhurst, G. (1978) ‘Electrochemical behavior of natural and biosynthetic

polynucleotides at the pyrolytic graphite electrode. A new probe for studies of polynucleotide

structure and reactions’, Journal of Electroanalytical Chemistry, 89, pp. 161–173.

Brabec, V. and Dryhurst, G. (1978b) ‘Electrochemical Oxidation of Polyadenylic-Acid at

Graphite Electrodes’, Journal of Electroanalytical Chemistry, 91(2), pp. 219–229. doi: Doi

10.1016/S0022-0728(78)80102-8.

Brabec, V. and Paleček, E. (1972) ‘Adsorption of single‐ stranded and double‐ helical

polynucleotides on the mercury electrode’, Biopolymers, 11(12), pp. 2577–2589. doi:

10.1002/bip.1972.360111215.

55

Brázda, V. et al. (2006) ‘Searching for target sequences by p53 protein is influenced by DNA

length’, Biochemical and Biophysical Research Communications, 341(2), pp. 470–477. doi:

10.1016/j.bbrc.2005.12.202.

Brázdilová, P. et al. (2007) ‘Ferrocenylethynyl Derivatives of Nucleoside Triphosphates:

Synthesis, Incorporation, Electrochemistry, and Bioanalytical Applications’, Chemistry - A

European Journal. WILEY‐ VCH Verlag, 13(34), pp. 9527–9533. doi:

10.1002/chem.200701249.

Brázdová, M. et al. (2002) ‘Role of tumor suppressor p53 domains in selective binding to

supercoiled DNA.’, Nucleic acids research, 30(22), pp. 4966–74. Available at:

http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=137164&tool=pmcentrez&rendert

ype=abstract.

Brázdová, M. et al. (2009) ‘Modulation of gene expression in U251 glioblastoma cells by

binding of mutant p53 R273H to intronic and intergenic sequences’, Nucleic Acids Research,

37(5), pp. 1486–1500. doi: 10.1093/nar/gkn1085.

Brázdová, M. et al. (2016) ‘P53 specifically binds triplex DNA in vitro and in cells’, PLoS

ONE, 11(12), pp. 1–25. doi: 10.1371/journal.pone.0167439.

Brotons, A. et al. (2016) ‘Carbon materials for the electrooxidation of nucleobases,

nucleosides and nucleotides toward cytosine methylation detection: a review’, Analytical

Methods, 8(4), pp. 702–715. doi: 10.1039/c5ay02616d.

Brugarolas, J. et al. (1995) ‘Radiation-induced cell cycle arrest compromised by p21

deficiency’, Nature, 377, pp. 552–557.

Cadwell, C. and Zambetti, G. P. (2001) ‘The effects of wild-type p53 tumor suppressor

activity and mutant p53 gain-of-function on cell growth.’, Gene, 277(1–2), pp. 15–30. doi:

Doi 10.1016/S0378-1119(01)00696-5.

Cahová, H. et al. (2008) ‘Aminophenyl- and nitrophenyl-labeled nucleoside triphosphates:

Synthesis, enzymatic incorporation, and electrochemical detection’, Angewandte Chemie -

International Edition, 47(11), pp. 2059–2062. doi: 10.1002/anie.200705088.

Čapek, P. et al. (2007) ‘An Efficient Method for the Construction of Functionalized DNA

Bearing Amino Acid Groups through Cross-Coupling Reactions of Nucleoside Triphosphates

Followed by Primer Extension or PCR’, Chemistry - A European Journal. Wiley-Blackwell,

13(21), pp. 6196–6203. doi: 10.1002/chem.200700220.

Chang, M. S. and Bollum, F. J. (1990) ‘Multiple Roles of Divalent

Deoxynucleotidyltransferase Cation in the Terminal Reaction *’, The Journal of Biological

Chemistry, 265(29), pp. 17436–17440.

Cheng, G. et al. (2010) ‘A new electrochemically active-inactive switching aptamer

molecular beacon to detect thrombin directly in solution’, Biosensors and Bioelectronics.

Elsevier B.V., 25(10), pp. 2265–2269. doi: 10.1016/j.bios.2010.03.008.

Christova, R. (2013) Detecting DNA-protein interactions in living cells - ChIP approach. 1st

edn, Advances in Protein Chemistry and Structural Biology. 1st edn. Elsevier Inc. doi:

10.1016/B978-0-12-411637-5.00004-4.

56

Cummings, T. E. and Elving, P. J. (1978) ‘Electrochemical Reduction of Uracil in Dimethyl-

Sulfoxide - Autoprotonation of Anion Radical’, Journal of Electroanalytical Chemistry,

94(2), pp. 123–145. doi: Doi 10.1016/S0022-0728(78)80329-5.

Cummings, T. E. and Elving, P. J. (1979) ‘Electrochemical Reduction of Thymine in

Dimethyl-Sulfoxide Auto-Protonation of the Radical-Anion and Reduced Free-Radical’,

Journal of Electroanalytical Chemistry, 102(2), pp. 237–248.

Dadová, J. et al. (2013) ‘Aqueous heck cross-coupling preparation of acrylate-modified

nucleotides and nucleoside triphosphates for polymerase synthesis of acrylate-labeled DNA’,

Journal of Organic Chemistry, 78(19). doi: 10.1021/jo4011574.

Daňhel, A., Havran, L., et al. (2016) ‘Hydrogen Evolution Facilitates Reduction of DNA

Guanine Residues at the Hanging Mercury Drop Electrode: Evidence for a Chemical

Mechanism’, Electroanalysis, 28(11), pp. 2785–2790. doi: 10.1002/elan.201600242.

Daňhel, A., Trošanová, Z., et al. (2016) ‘Voltammetric analysis of 5-(4-Azidophenyl)-2’-

deoxycytidine nucleoside and azidophenyl-labelled single- and double-stranded DNAs’,

Electrochimica Acta, 215(August), pp. 72–83. doi: 10.1016/j.electacta.2016.08.096.

Deibel, M. R. and Coleman, M. S. (1980) ‘Biochemical properties of purified human terminal

deoxynucleotidyltransferase.’, The Journal of biological chemistry, 255(9), pp. 4206–12.

Available at: http://www.ncbi.nlm.nih.gov/pubmed/7372675.

Deng, C. et al. (1995) ‘Mice lacking p21CIP1/WAF1 undergo normal development, but are

defective in G1 checkpoint control.’, Cell, 82(4), pp. 675–684. doi: 10.1016/0092-

8674(95)90039-X.

Deng, G. and Wu, R. (1981) ‘An improved procedure for utilizing terminal transferase to add

homopolymers to the 3’ termini of DNA’, Nucleic Acids Research, 9(16), pp. 4173–4188.

DeRosa, M. C., Sancar, A. and Barton, J. K. (2005) ‘Electrically monitoring DNA repair by

photolyase.’, Proceedings of the National Academy of Sciences of the United States of

America, 102(31), pp. 10788–92. doi: 10.1073/pnas.0503527102.

Deubel, D. V. (2003) ‘Reactivity of osmium tetraoxide towards nitrogen heterocycles:

Implications for the molecular recognition of DNA mismatch’, Angewandte Chemie -

International Edition, 42(17), pp. 1974–1977. doi: 10.1002/anie.200250462.

Dey, B. et al. (2012) ‘DNA-protein interactions: Methods for detection and analysis’,

Molecular and Cellular Biochemistry, 365(1–2), pp. 279–299. doi: 10.1007/s11010-012-

1269-z.

Ding, C., Ge, Y. and Lin, J. M. (2010) ‘Aptamer based electrochemical assay for the

determination of thrombin by using the amplification of the nanoparticles’, Biosensors and

Bioelectronics, 25(6), pp. 1290–1294. doi: 10.1016/j.bios.2009.10.017.

Dötsch, V. et al. (2010) ‘P63 and P73, the Ancestors of P53.’, Cold Spring Harbor

perspectives in biology, 2(9). doi: 10.1101/cshperspect.a004887.

Dryhurst, G. and Pace, G. F. (1970) ‘Electrochemical Oxidation of Guanine at the Pyrolytic

Graphite Electrode’, Journal of The Electrochemical Society, 117(10), p. 1259. doi:

10.1149/1.2407283.

57

Dudová, Z. et al. (2016) ‘Electrochemical behavior of 7-deazaguanine- and 7-deazaadenine-

modified DNA at the hanging mercury drop electrode’, Monatshefte fur Chemie, 147(1), pp.

3–11. doi: 10.1007/s00706-015-1584-7.

El-Deiry, W. S. et al. (1992) ‘Definition of a consensus binding site for p53’, Nature

Genetics. Nature Publishing Group, 1(1), pp. 45–49. doi: 10.1038/ng0492-45.

Evtugyn, G. A. et al. (2008) ‘Potentiometric DNA sensor based on electropolymerized

phenothiazines for protein detection’, Electroanalysis, 20(12), pp. 1300–1308. doi:

10.1002/elan.200704186.

Fadrná, R. et al. (2004) ‘Voltammetric Determination of Adenine, Guanine, and DNA Using

Liquid Mercury Free Polished Silver Solid Amalgam Electrode’, Analytical Letters, 37(3), pp.

399–413. doi: 10.1081/AL-120028615.

Ferapontova, E. E. and Domínguez, E. (2003) ‘Direct electrochemical oxidation of DNA on

polycrystalline gold electrodes’, Electroanalysis, 15(7), pp. 629–634. doi:

10.1002/elan.200390079.

Fields, S. and Jang, S. K. (1990) ‘Presence of a Potent Transcription Activating Sequence in

the p53 Protein’, Science, 249(4972), pp. 1046–1049. doi: 10.1126/science.2144363.

Flechsig, G.-U. and Reske, T. (2007) ‘Electrochemical detection of DNA hybridization by

means of osmium tetroxide complexes and protective oligonucleotides’, Analytical

Chemistry. American Chemical Society, 79(5), pp. 2125–2130. doi: 10.1021/ac062075c.

Fojta, M. et al. (1998) ‘Two superhelix density-dependent DNA transitions detected by

changes in DNA adsorption/desorption behavior’, Biochemistry, 37(14), pp. 4853–4862. doi:

10.1021/bi9729559.

Fojta, M. (2002) ‘Electrochemical Sensors for DNA Interactions and Damage’,

Electroanalysis, 14(21), pp. 1449–1463.

Fojta, M. et al. (2002) ‘Voltammetric microanalysis of DNA adducts with osmium

tetroxide,2,2’-bipyridine using a pyrolytic graphite electrode’, Talanta, 56(5), pp. 867–874.

Fojta, M., Pivonkova, H., et al. (2003) ‘Recognition of DNA modified by antitumor cisplatin

by “latent” and “active” protein p53’, Biochemical Pharmacology, 65(8), pp. 1305–1316. doi:

10.1016/S0006-2952(03)00078-9.

Fojta, M., Havran, L., et al. (2003) ‘Two-surface strategy in electrochemical DNA

hybridization assays: Detection of osmium-labeled target DNA at carbon electrodes’,

Electroanalysis, 15(5–6), pp. 431–440. doi: 10.1002/elan.200390050.

Fojta, M., Pivonkova, H., et al. (2004) ‘Investigations of the supercoil-selective DNA binding

of wild type p53 suggest a novel mechanism for controlling p53 function’, European Journal

of Biochemistry, 271(19), pp. 3865–3876. doi: 10.1111/j.1432-1033.2004.04323.x.

Fojta, M. (2004) ‘Mercury Electrodes in Nucleic Acid Electrochemistry: Sensitive Analytical

Tools and Probes of DNA Structure. A Review’, Collection of Czechoslovak Chemical

Communications, 69(4), pp. 715–747. doi: 10.1135/cccc20040715.

Fojta, M., Havran, L., et al. (2004) ‘Multiply osmium-labeled reporter probes for

58

electrochemical DNA hybridization assays: Detection of trinucleotide repeats’, Biosensors

and Bioelectronics, 20(5), pp. 985–994. doi: 10.1016/j.bios.2004.06.015.

Fojta, M. (2005) ‘Use of DNA Repair Enzymes in Electrochemical Detection of Damage to

DNA Bases in Vitro and in’, Bioelectrochemistry, 77(9), pp. 2920–2927.

Fojta, M. et al. (2007) ‘“Multicolor” electrochemical labeling of DNA hybridization probes

with osmium tetroxide complexes’, Analytical Chemistry, 79(3), pp. 1022–1029. doi:

10.1021/ac0616299.

Fojta, M. et al. (2008) ‘Electrochemical stripping techniques in analysis of nucleic acids and

their constituents’, Current Analytical Chemistry, 4(3), pp. 250–262. doi:

10.2174/157341108784911415.

Fojta, M. et al. (2011) ‘Osmium tetroxide complexes as versatile tools for structure probing

and electrochemical analysis of biopolymers’, Current Analytical Chemistry, 7(1), pp. 35–50.

doi: 10.2174/157341111793797653.

Fojta, M. et al. (2016) ‘Recent progress in electrochemical sensors and assays for DNA

damage and repair’, TrAC - Trends in Analytical Chemistry, 79, pp. 160–167. doi:

10.1016/j.trac.2015.11.018.

Fojta, M. and Paleček, E. (1997) ‘Supercoiled DNA-modified mercury electrode: A highly

sensitive tool for the detection of DNA damage’, Analytica Chimica Acta, 342(1), pp. 1–12.

doi: 10.1016/s0003-2670(96)00551-x.

Foord, O., Navot, N. and Rotter, V. (1993) ‘Isolation and characterization of DNA sequences

that are specificially bound by wild-type p53 protein’, Molecular and Celluar Biology, 13, pp.

1378–1384.

Furey, T. S. (2012) ‘ChIP-seq and Beyond: new and improved methodologies to detect and

characterize protein-DNA interactions’, Nature Reviews Genetics, 13(12), pp. 840–852. doi:

10.1038/nrg3306.ChIP-seq.

Gebala, M. et al. (2009) ‘Label-Free Detection of DNA Hybridization in Presence of

Intercalators Using Electrochemical Impedance Spectroscopy’, Electroanalysis. Wiley-

Blackwell, 21(3–5), pp. 325–331. doi: 10.1002/elan.200804388.

Giono, L. E. and Manfredi, J. J. (2006) ‘The p53 Tumor Suppressor Participates in Multiple

Cell Cycle Checkpoints’, Journal of cellular physiology, 209(1), pp. 13–20. doi: 10.1002/JCP.

Göhler, T. et al. (2002) ‘Specific interaction of p53 with target binding sites is determined by

DNA conformation and is regulated by the C-terminal domain’, Journal of Biological

Chemistry, 277(43), pp. 41192–41203. doi: 10.1074/jbc.M202344200.

Gouge, J. et al. (2013) ‘Structures of intermediates along the catalytic cycle of terminal

deoxynucleotidyltransferase: Dynamical aspects of the two-metal ion mechanism’, Journal of

Molecular Biology. Elsevier Ltd, 425(22), pp. 4334–4352. doi: 10.1016/j.jmb.2013.07.009.

Green, D. R. and Kroemer, G. (2009) ‘Cytoplasmic functions of the tumor suppressor p53’,

Nature, 458(7242), p. 1127. doi: 10.1038/nature07986.Cytoplasmic.

Halazonetis, T. D., Davis, L. J. and Kandil, A. N. (1993) ‘Wild-type p53 adopts a ’mutant’-

59

like conformation when bound to DNA.’, The EMBO journal, 12(3), pp. 1021–8. doi:

10.1002/j.1460-2075.1993.tb05743.x.

Hasoň, S. and Vetterl, V. (2006) ‘Microanalysis of oligodeoxynucleotides by cathodic

stripping voltammetry at amalgam-alloy surfaces in the presence of copper ions’, Talanta,

69(3), pp. 572–580. doi: 10.1016/j.talanta.2005.10.029.

Hasoň, S., Vetterl, V. and Fojta, M. (2008) ‘Two-dimensional condensation of pyrimidine

oligonucleotides during their self-assemblies at mercury based surfaces’, Electrochimica Acta,

53(6), pp. 2818–2824. doi: 10.1016/j.electacta.2007.10.065.

Havran, L. et al. (2008) ‘Sensitive voltammetric detection of DNA damage at carbon

electrodes using DNA repair enzymes and an electroactive osmium marker’, Analytical and

Bioanalytical Chemistry, 391(5), pp. 1751–1758. doi: 10.1007/s00216-008-1850-1.

Havranová-Vidláková, P. et al. (2017) ‘Butylacrylate-nucleobase Conjugates as Targets for

Two-step Redox Labeling of DNA with an Osmium Tetroxide Complex’, Electroanalysis, pp.

371–377. doi: 10.1002/elan.201700702.

Havranová-Vidláková, P. et al. (2018) ‘Butylacrylate-nucleobase Conjugates as Targets for

Two-step Redox Labeling of DNA with an Osmium Tetroxide Complex’, Electroanalysis,

30(2). doi: 10.1002/elan.201700702.

Herne, T. M. and Tarlov, M. J. (1997) ‘Characterization of DNA Probes Immobilized on Gold

Surfaces’, J. Am. Chem. Soc., 119(13), pp. 8916–8920. doi: 10.1021/ja9719586.

Hippel, P. H. von (1994) ‘Protein-DNA Recognition: New Perspectives and Underlying

Themes’, Science, 263, pp. 769–770.

Hocek, M. and Fojta, M. (2008) ‘Cross-coupling reactions of nucleoside triphosphates

followed by polymerase incorporation. Construction and applications of base-functionalized

nucleic acids’, Org Biomol Chem, 6(13), pp. 2233–2241. doi: 10.1039/b803664k.

Hocek, M. and Fojta, M. (2011) ‘Nucleobase modification as redox DNA labelling for

electrochemical detection’, Chemical Society Reviews, 40(12), p. 5802. doi:

10.1039/c1cs15049a.

Horakova, P. et al. (2011) ‘Tail-labelling of DNA probes using modified deoxynucleotide

triphosphates and terminal deoxynucleotidyl transferase. Application in electrochemical DNA

hybridization and protein-DNA binding assays’, Org Biomol Chem. 2011/01/05, 9(5), pp.

1366–1371. doi: 10.1039/c0ob00856g.

Horáková, P. et al. (2011) ‘Tail-labelling of DNA probes using modified deoxynucleotide

triphosphates and terminal deoxynucleotidyl transferase. Application in electrochemical DNA

hybridization and protein-DNA binding assays.’, Organic & biomolecular chemistry, 9(5),

pp. 1366–71. doi: 10.1039/c0ob00856g.

Huang, H. et al. (2009) ‘DNA aptamer-based detection of lysozyme by an

electrochemiluminescence assay coupled to quantum dots’, Electrochemistry

Communications. Elsevier B.V., 11(4), pp. 816–818. doi: 10.1016/j.elecom.2009.01.009.

Hupp, T. R. (1999) ‘Regulation of p53 protein function through alterations in protein-folding

pathways’, Cellular and Molecular Life Sciences, 55(1), pp. 88–95. doi:

60

10.1007/s000180050272.

Hupp, T. R. and Lane, D. P. (1994) ‘Regulation of the cryptic sequence-specific DNA-

binding function of p53 by protein kinases’, Cold Spring Harbor Symposia on Quantitative

Biology, 59, pp. 195–206. doi: 10.1101/SQB.1994.059.01.024.

Jacobsen, M. and Flechsig, G.-U. (2013) ‘Hybridization Detection of Osmium Tetroxide

Bipyridine-Labeled DNA and RNA on Heated Gold Wire Electrodes’, Electroanalysis.

Wiley-Blackwell, 25(2), pp. 373–379. doi: 10.1002/elan.201200460.

Jagelská, E. B. et al. (2010) ‘The potential of the cruciform structure formation as an

important factor influencing p53 sequence-specific binding to natural DNA targets’,

Biochemical and Biophysical Research Communications, 391(3), pp. 1409–1414. doi:

10.1016/j.bbrc.2009.12.076.

Jäger, S. et al. (2005) ‘A versatile toolbox for variable DNA functionalization at high

density’, Journal of the American Chemical Society, 127(43), pp. 15071–15082. doi:

10.1021/ja051725b.

Jelen, F. et al. (1991) ‘Osmium tetroxide reactivity of DNA bases in nucleotide sequencing

and probing of DNA structure’, General Physiology and Biophysics, 10(5), pp. 461–473.

Jelen, F., Erdem, A. and Paleček, E. (2002) ‘Cyclic voltammetry of echinomycin and its

interaction with double-stranded and single-stranded DNA adsorbed at the electrode’,

Bioelectrochemistry, 55(1–2), pp. 165–167. doi: 10.1016/S1567-5394(01)00143-8.

Jen-Jacobson, L. (1997) ‘Protein-DNA recognition complexes: Conservation of structure and

binding energy in the transition state’, Biopolymers - Nucleic Acid Sciences Section, 44(2), pp.

153–180. doi: 10.1002/(SICI)1097-0282(1997)44:2<153::AID-BIP4>3.0.CO;2-U.

Kaelin, W. G. (1999) ‘The p53 gene family.’, Oncogene, 18(53), pp. 7701–7705. doi:

10.1038/sj.onc.1202955.

Kato, K.-I. et al. (1967) ‘Deoxynucleotide-polymerizing Thymus Gland Enzymes of Calf’,

The Journal of Biological Chemistry, 242(11), pp. 2780–2789.

Kawde, A. N. et al. (2005) ‘Label-free bioelectronic detection of aptamer-protein

interactions’, Electrochemistry Communications, 7(5), pp. 537–540. doi:

10.1016/j.elecom.2005.03.008.

Khoury, M. P. and Bourdon, J. (2010) ‘The Isoforms of the p53 Protein’, Cold Spring Harbor

perspectives in biology, pp. 1–11. doi: 10.1101/cshperspect.a000927.

Kim, E. and Deppert, W. (2003) ‘The complex interactions of p53 with target DNA : we learn

as we go’, Cell, 150, pp. 141–150. doi: 10.1139/O03-046.

Komori, T. et al. (1993) ‘Lack of N regions in antigen receptor variable region genes of TdT-

deficient lymphocytes’, Science, 261(5125), pp. 1171–1175. doi: 10.1126/science.8356451.

Kubbutat, M. H. G., Jones, S. N. and Vousden, K. H. (1997) ‘Regulation of p53 stability by

Mdm2’, Nature, 387, pp. 299–303.

Kunkel, T. a et al. (1986) ‘Rearrangements of DNA mediated by terminal transferase.’,

Proceedings of the National Academy of Sciences of the United States of America, 83(6), pp.

61

1867–71. doi: 10.1073/pnas.83.6.1867.

Lamb, P. and Crawford, L. (1986) ‘Characterization of the Human p53 Gene’, Molecular and

Cellular Biology, 6(5), pp. 1379–1385.

Lane, D. P. (1992) ‘Cancer. p53, guardian of the genome.’, Nature, pp. 15–16. doi:

10.1038/358015a0.

Lane, D. P. and Crawford, L. V (1979) ‘T antigen is bound to a host protein in SV40-

transformed cells.’, Nature, 278, pp. 261–263. doi: 10.1038/278261a0.

Langer, P. R., Waldrop, A. A. and Ward, D. C. (1981) ‘Enzymatic synthesis of biotin-labeled

polynucleotides: novel nucleic acid affinity probes.’, Proceedings of the National Academy of

Sciences, 78(11), pp. 6633–6637. doi: 10.1073/pnas.78.11.6633.

Latonen, L. and Laiho, M. (2005) ‘Cellular UV damage responses - Functions of tumor

suppressor p53’, Biochimica et Biophysica Acta - Reviews on Cancer, 1755(2), pp. 71–89.

doi: 10.1016/j.bbcan.2005.04.003.

Levine, A. J. (1997) ‘P53, the Cellular Gatekeeper for Growth and Division’, Cell, 88(3), pp.

323–331. doi: 10.1016/S0092-8674(00)81871-1.

Liu, F. et al. (2015) ‘An electrochemical method to assay human 8-oxoguanine DNA

glycosylase 1’, Electrochemistry Communications. Elsevier B.V., 50, pp. 51–54. doi:

10.1016/j.elecom.2014.11.011.

Masařík, M. et al. (2007) ‘Label-free voltammetric detection of single-nucleotide mismatches

recognized by the protein MutS’, Analytical and Bioanalytical Chemistry, 388(1), pp. 259–

270. doi: 10.1007/s00216-007-1181-7.

Mazur, S. J. et al. (1999) ‘Preferential binding of tumor suppressor p53 to positively or

negatively supercoiled DNA involves the C-terminal domain.’, Journal of molecular biology,

292(2), pp. 241–249. doi: 10.1006/jmbi.1999.3064.

McKinney, K. and Prives, C. (2002) ‘Efficient specific DNA binding by p53 requires both its

central and C-terminal domains as revealed by studies with high-mobility group 1 protein.’,

Molecular and cellular biology, 22(19), pp. 6797–808. doi: 10.1128/MCB.22.19.6797.

Menova, P., Raindlova, V. and Hocek, M. (2013) ‘Scope and Limitations of the Nicking

Enzyme Amplification Reaction for the Synthesis of Base-Modified Oligonucleotides and

Primers for PCR’, Bioconjugate Chemistry, 24(6), pp. 1081–1093. doi: 10.1021/bc400149q.

Meunier-Prest, R. et al. (2010) ‘Electrochemical probe for the monitoring of DNA-protein

interactions’, Biosensors and Bioelectronics. Elsevier B.V., 25(12), pp. 2598–2602. doi:

10.1016/j.bios.2010.04.023.

Michael, D. and Oren, M. (2003) ‘The p53-Mdm2 module and the ubiquitin system’,

Seminars in Cancer Biology. Elsevier, 13(1), pp. 49–58. doi: 10.1016/S1044-579X(02)00099-

8.

Michelson, A. M. and Orkin, S. H. (1982) ‘Characterization of the homopolymer tailing

reaction catalyzed by terminal deoxynucleotidyl transferase. Implications for the cloning of

cDNA.’, Journal of Biological Chemistry, 257(24), pp. 14773–14782.

62

Mickelsen, S. et al. (1999) ‘Modulation of terminal deoxynucleotidyltransferase activity by

the DNA-dependent protein kinase.’, Journal of immunology, 163(2), pp. 834–843.

Mikelová, R. et al. (2007) ‘Resolution of overlapped reduction signals in short hetero-

oligonucleotides by elimination voltammetry’, Electroanalysis, 19(2–3), pp. 348–355. doi:

10.1002/elan.200603739.

Modak, M. J. (1978) ‘Biochemistry of Terminal Deoxynucleotidyltransferase: Mechanism of

Inhibition by Adenosine 5ʹ-Triphosphate’, Biochemistry, 17(15), pp. 3116–3120. doi:

10.1021/bi00608a027.

Motea, E. A. and Berdis, A. J. (2010) ‘Terminal deoxynucleotidyl transferase: The story of a

misguided DNA polymerase’, Biochimica et Biophysica Acta - Proteins and Proteomics.

Elsevier B.V., 1804(5), pp. 1151–1166. doi: 10.1016/j.bbapap.2009.06.030.

Němcová, K. et al. (2010) ‘A label-free electrochemical test for DNA-binding activities of

tumor suppressor protein p53 using immunoprecipitation at magnetic beads’, Analytica

Chimica Acta, 668(2), pp. 166–170. doi: 10.1016/j.aca.2010.04.018.

Němcová, K. et al. (2014) ‘Electrochemical detection of DNA binding by tumor suppressor

p53 protein using osmium-labeled oligonucleotide probes and catalytic hydrogen evolution at

the mercury electrode’, Analytical and Bioanalytical Chemistry, 406(24), pp. 5843–5852. doi:

10.1007/s00216-014-7996-0.

Oda, E. (2000) ‘Noxa, a BH3-Only Member of the Bcl-2 Family and Candidate Mediator of

p53-Induced Apoptosis’, Science, 288(5468), pp. 1053–1058. doi:

10.1126/science.288.5468.1053.

Oliveira-Brett, A. M. et al. (2004) ‘Voltammetric determination of all DNA nucleotides’,

Anal Biochem, 332(2), pp. 321–329. doi: 10.1016/j.ab.2004.06.021.

Ono, K. (1990) ‘Inhibitory effects of various 2′, 3′-dideoxynucleoside 5′-triphosphates on the

utilization of 2′-deoxynucleoside 5′-triphosphates by terminal deoxynucleotidyltransferase

from calf thymus’, BBA - Gene Structure and Expression, 1049(1), pp. 15–20. doi:

10.1016/0167-4781(90)90078-G.

Oren, M. (1999) ‘Regulation of the p53 Tumor Suppressor Protein’, Journal of Biological

Chemistry, 274(51), pp. 36031–36034. doi: 10.1074/jbc.274.51.36031.

Ostatná, V. et al. (2017) ‘Chronopotentiometric sensing of specific interactions between

lysozyme and the DNA aptamer’, Bioelectrochemistry, 114, pp. 42–47. doi:

10.1016/j.bioelechem.2016.12.003.

Owen-Schaub, L. B. et al. (1995) ‘Wild-type human p53 and a temperature-sensitive mutant

induce Fas/APO-1 expression.’, Molecular and Cellular Biology, 15(6), pp. 3032–3040. doi:

10.1128/MCB.15.6.3032.

Palecek, E. et al. (2001) ‘Binding of p53 and its core domain to supercoiled DNA’, European

Journal of Biochemistry, 268(3), pp. 573–581. doi: 10.1046/j.1432-1327.2001.01898.x.

Paleček, E. (1958) ‘Oszillographische Polarographie der Nucleeinsäuren und ihrer

Bestandteile’, Naturwissenchaften, 45(8), pp. 186–187.

63

Paleček, E. (1960) ‘Oscillographic polarography of highly polymerized Deoxyribonucleic

acid’, Nature, 188, pp. 656–657. doi: 10.1038/188656a0.

Paleček, E. (1992a) ‘Probing DNA structure with osmium tetroxide complexes in vitro’,

Methods in Enzymology. Academic Press, 212(C), pp. 139–155. doi: 10.1016/0076-

6879(92)12010-N.

Paleček, E. (1992b) ‘Probing DNA structure with osmium tetroxide complexes in vitro’,

Methods in Enzymology, 212, pp. 139–155. doi: 10.1016/0076-6879(92)12010-N.

Paleček, E. et al. (1997) ‘Tumor suppressor protein p53 binds preferentially to supercoiled

DNA’, Oncogene, 15(18), pp. 2201–2209. doi: 10.1038/sj.onc.1201398.

Paleček, E. et al. (2004) ‘Sensitive electrochemical determination of unlabeled MutS protein

and detection of point mutations in DNA.’, Analytical chemistry, 76(19), pp. 5930–5936. doi:

10.1021/ac049474x.

Paleček, E. and Bartošík, M. (2012) ‘Electrochemistry of nucleic acids’, Chemical Reviews,

112(6), pp. 3427–3481. doi: 10.1021/cr200303p.

Palecek, E. and Fojta, M. (2005) ‘Electrochemical DNA Sensors’, in Bioelectronics.

Weinheim, FRG: Wiley-VCH Verlag GmbH & Co. KGaA, pp. 127–192. doi:

10.1002/352760376X.ch5.

Paleček, E. and Fojta, M. (2007) ‘Magnetic beads as versatile tools for electrochemical DNA

and protein biosensing’, Talanta, 74(3), pp. 276–290. doi: 10.1016/j.talanta.2007.08.020.

Palecek, E. and Hung, M. A. (1983) ‘Determination of nanogram quantities of osmium-

labeled nucleic acids by stripping (inverse) voltammetry.’, Analytical biochemistry, 132(2),

pp. 236–42. Available at: http://www.ncbi.nlm.nih.gov/pubmed/6353997 (Accessed: 5 April

2017).

Paleček, E. and Postbieglová, I. (1986) ‘Adsorptive stripping voltammetry of

biomacromolecules with transfer of the adsorbed layer’, Journal of Electroanalytical

Chemistry, 214(1–2), pp. 359–371. doi: 10.1016/0022-0728(86)80108-5.

Paleček, E., Trefulka, M. and Fojta, M. (2009) ‘End-labeling of peptide nucleic acid with

osmium complex. Voltammetry at carbon and mercury electrodes’, Electrochemistry

Communications, 11(2), pp. 359–362. doi: 10.1016/j.elecom.2008.11.049.

Pivoňková, H. et al. (2006) ‘Recognition of cisplatin-damaged DNA by p53 protein: Critical

role of the p53 C-terminal domain’, Biochemical and Biophysical Research Communications,

339(2), pp. 477–484. doi: 10.1016/j.bbrc.2005.11.038.

Pivoňková, H. et al. (2010) ‘Direct voltammetric analysis of DNA modified with

enzymatically incorporated 7-deazapurines’, Analytical Chemistry, 82(16), pp. 6807–6813.

doi: 10.1021/ac100757v.

Polyak, K. et al. (1997) ‘A model for p53-induced apoptosis’, Nature, 389(6648), pp. 300–

305. doi: 10.1038/38525.

Ramadan, K. et al. (2004) ‘De Novo DNA synthesis by human DNA polymerase λ, DNA

polymerase μ and terminal deoxyribonucleotidyl transferase’, Journal of Molecular Biology,

64

339(2), pp. 395–404. doi: 10.1016/j.jmb.2004.03.056.

Reske, T. et al. (2009) ‘Kinetics of the labeling reactions of thymine, cytosine and uracil with

osmium tetroxide bipyridine’, Microchimica Acta. Springer Vienna, 166(3–4), pp. 197–201.

doi: 10.1007/s00604-009-0195-6.

Riedl, J. et al. (2009) ‘Tetrathiafulvalene-Labelled Nucleosides and Nucleoside

Triphosphates: Synthesis, Electrochemistry and the Scope of Their Polymerase Incorporation

into DNA’, European Journal of Organic Chemistry, 21, pp. 3519–3525. doi:

10.1002/ejoc.200900392.

Roychoudhury, R., Jay, E. and Wu, R. (1976) ‘Terminal labeling and addition of

homopolymer tracts to duplex DNA fragments by terminal deoxynucleotidyl transferase.’,

Nucleic acids research, 3(4), pp. 863–77. doi: 10.1093/nar/3.1.101.

Sadofsky, M. J. (2001) ‘The RAG proteins in V(D)J recombination: more than just a

nuclease’, Nucleic Acids Research, 29(7), pp. 1399–1409. doi: 10.1093/nar/29.7.1399.

Sakamuro, D. et al. (1997) ‘The polyproline region of p53 is required to activate apoptosis but

not growth arrest.’, Oncogene, 15(8), pp. 887–898. doi: 10.1038/sj.onc.1201263.

Seeman, N. C., Rosenberg, J. M. and Rich, A. (1976) ‘Sequence-specific recognition of

double helical nucleic acids by proteins.’, Proceedings of the National Academy of Sciences,

73(3), pp. 804–808. doi: 10.1073/pnas.73.3.804.

Selivanova, G. et al. (1998) ‘Reactivation of mutant p53: a new strategy for cancer therapy.’,

Seminars in cancer biology, 8(5), pp. 369–378. doi: 10.1006/scbi.1998.0099.

Simonova, A. et al. (2014) ‘Methoxyphenol and Dihydrobenzofuran as Oxidizable Labels for

Electrochemical Detection of DNA’, ChemPlusChem. WILEY‐ VCH Verlag, 79(12), p. n/a-

n/a. doi: 10.1002/cplu.201402194.

Simonova, A. et al. (2017) ‘Phenothiazine-linked Nucleosides and Nucleotides for Redox

Labelling of DNA’, Org. Biomol. Chem., (July). doi: 10.1039/C7OB01439B.

Smith, M. L. et al. (2000) ‘p53-mediated DNA repair responses to UV radiation: studies of

mouse cells lacking p53, p21, and/or gadd45 genes.’, Molecular and cellular biology, 20(10),

pp. 3705–3714. doi: 10.1128/MCB.20.10.3705-3714.2000.

Špaček, J. et al. (2017) ‘Label-free detection of canonical DNA bases, uracil and 5-

methylcytosine in DNA oligonucleotides using linear sweep voltammetry at a pyrolytic

graphite electrode’, Electrochemistry Communications, 82(June), pp. 34–38. doi:

10.1016/j.elecom.2017.07.013.

Stansel, R. M., Subramanian, D. and Griffith, J. D. (2002) ‘P53 Binds Telomeric Single

Strand Overhangs and T-Loop Junctions in Vitro’, Journal of Biological Chemistry, 277(14),

pp. 11625–11628. doi: 10.1074/jbc.C100764200.

Studničková, M. et al. (1989) ‘Reduction of guanosine at a mercury electrode’,

Bioelectrochemistry and Bioenergetics, 275(21), pp. 83–86.

Subramanian, D. and Griffith, J. D. (2005) ‘Modulation of p53 binding to Holliday junctions

and 3-cytosine bulges by phosphorylation events’, Biochemistry, 44(7), pp. 2536–2544. doi:

65

10.1021/bi048700u.

Surkus, A. E. and Flechsig, G.-U. (2009) ‘Electrochemical detection of DNA melting curves

by means of heated biosensors’, Electroanalysis, 21(10), pp. 1119–1123. doi:

10.1002/elan.200804539.

Toshiyuki, M. and Reed, J. C. (1995) ‘Tumor suppressor p53 is a direct transcriptional

activator of the human bax gene’, Cell, 80(2), pp. 293–299. doi: 10.1016/0092-

8674(95)90412-3.

Trefulka, M. et al. (2007) ‘Voltammetry of osmium end-labeled oligodeoxynucleotides at

carbon, mercury, and gold electrodes’, Electroanalysis. Wiley-Blackwell, 19(12), pp. 1334–

1338. doi: 10.1002/elan.200703849.

Trefulka, M., Bartošík, M. and Paleček, E. (2010) ‘Facile end-labeling of RNA with

electroactive Os(VI) complexes’, Electrochemistry Communications. Elsevier, 12(12), pp.

1760–1763. doi: 10.1016/J.ELECOM.2010.10.016.

Trnková, L., Jelen, F. and Postbieglová, I. (2003) ‘Application of Elimination Voltammetry to

the Resolution of Adenine and Cytosine Signals in Oligonucleotides. I. Homo-

oligodeoxynucleotides dA9 and dC9’, Electroanalysis, 15(19), pp. 1529–1535. doi:

10.1002/elan.200302719.

Trnková, L., Jelen, F. and Postbieglová, I. (2006) ‘Application of elimination voltammetry to

the resolution of adenine and cytosine signals in oligonucleotides II. Hetero-

oligodeoxynucleotides with different sequences of adenine and cytosine nucleotides’,

Electroanalysis, 18(7), pp. 662–669. doi: 10.1002/elan.200503447.

Trnková, L., Studničková, M. and Paleček, E. (1980) ‘Electrochemical behaviour of guanine

and its derivatives. I. Fast cyclic voltammetry of guanosine and calf thymus DNA’, Journal of

Electroanalytical Chemistry, 116(7), pp. 643–658. doi: 10.1016/S0022-0728(80)80294-4.

Vacek, J. et al. (2008) ‘Label-Free Electrochemical Monitoring of DNA Ligase Activity’,

Analytical Chemistry, 80(19), pp. 7609–7613.

Vidláková, P. et al. (2015) ‘G-quadruplex-based structural transitions in 15-mer DNA

oligonucleotides varying in lengths of internal oligo(dG) stretches detected by voltammetric

techniques’, Analytical and bioanalytical chemistry, 407(19), pp. 5817–5826. doi:

10.1007/s00216-015-8768-1.

Vojtĕsek, B. et al. (1992) ‘An immunochemical analysis of the human nuclear

phosphoprotein p53. New monoclonal antibodies and epitope mapping using recombinant

p53.’, Journal of immunological methods, 151(1–2), pp. 237–44. doi: 10.1016/0022-

1759(92)90122-A.

Vrábel, M. et al. (2009) ‘Base-modified DNA labeled by [Ru(bpy)3]2+ and [Os(bpy)3]2+

complexes: Construction by polymerase incorporation of modified nucleoside triphosphates,

electrochemical and luminescent properties, and applications’, Chemistry - A European

Journal, 15(5), pp. 1144–1154. doi: 10.1002/chem.200801538.

Wang, J. (2001a) Analytical Electrochemistry, Analytical Electrochemistry. doi:

10.1002/0471790303.ch3.

66

Wang, J. (2001b) Analytical Electrochemistry, Analytical Electrochemistry. doi:

10.1002/0471790303.ch3.

Wang, W. et al. (2005) ‘Acridine derivatives activate p53 and induce tumor cell death through

bax’, Cancer Biology and Therapy, 4(8), pp. 893–898. doi: 10.4161/cbt.4.8.2134.

Wang, Y. et al. (1995) ‘Interaction of p53 with its consensus DNA-binding site.’, Molecular

and Cellular Biology, 15(4), pp. 2157–2165. doi: 10.1128/MCB.15.4.2157.

Wong, E. L. S. and Gooding, J. J. (2003) ‘Electronic detection of target nucleic acids by a 2,6-

disulfonic acid anthraquinone intercalator’, Analytical Chemistry, 75(15), pp. 3845–3852. doi:

10.1021/ac034129d.

Wu, X. et al. (1993) ‘The p53-mdm-2 autoregulatory feedback loop’, Genes & Development,

53(7), pp. 1126–1132. doi: 10.1101/gad.7.7a.1126.

Yang, B., Gathy, K. N. and Coleman, M. S. (1994) ‘Mutational analysis of residues in the

nucleotide binding domain of human terminal deoxynucleotidyl transferase’, Journal of

Biological Chemistry, 269(16), pp. 11859–11868.

Yosypchuk, B. et al. (2002) ‘Use of solid amalgam electrodes in nucleic acid analysis’,

Electroanalysis, 14(21), pp. 1488–1493. doi: 10.1002/1521-4109(200211)14:21<1488::AID-

ELAN1488>3.0.CO;2-X.

Yosypchuk, B. et al. (2006) ‘Voltammetric behavior of osmium-labeled DNA at mercury

meniscus-modified solid amalgam electrodes. Detecting DNA hybridization’, Electroanalysis,

18(2), pp. 186–194. doi: 10.1002/elan.200503392.

Zauner, G. et al. (2005) ‘Tethered DNA hairpins facilitate electrochemical detection of DNA

ligation’, The Analyst. The Royal Society of Chemistry, 130(3), p. 345. doi:

10.1039/b413556c.

Zheng, J. et al. (2010) ‘An aptamer-based assay for thrombin via structure switch based on

gold nanoparticles and magnetic nanoparticles’, Talanta. Elsevier B.V., 80(5), pp. 1868–

1872. doi: 10.1016/j.talanta.2009.10.036.

Zhou, M., Zhai, Y. and Dong, S. (2009) ‘Electrochemical sensing and biosensing platform

based on chemically reduced graphene oxide’, Analytical Chemistry, 81(14), pp. 5603–5613.

doi: 10.1021/ac900136z.

Zilfou, J. T. and Lowe, S. W. (2009) ‘Tumor suppressive functions of p53.’, Cold Spring

Harbor perspectives in biology, 1(5). doi: 10.1101/cshperspect.a001883.

67

List of abbreviations

scDNA supercoiled DNA

ACV alternating current voltammetry

AdTSV adsorptive transfer stripping voltammetry

AFM atomic force microscopy

BA butylacrylate

BF benzofurazane

bpy 2,2’-bipyridine

ChIP chromatin immunoprecipitation

CON DNA probe containing p53CON

CTDBS C-terminal DNA binding site of the p53 protein

CV cyclic voltammetry

DME dropping mercury electrode

DPV differential-pulse voltammetry

dsDNA double-stranded DNA

EMSA electrophoretic mobility shift assay

FRET fluorescence resonance energy transfer

G4 G-quadruplexes

HMDE hanging mercury drop electrode

Km Michaelis constant

LSV linear sweep voltammetry

MB magnetic beads

68

MBIP magnetic beads-based immunoprecipitation assay

MFE mercury film electrode

NA nucleic acids

NEAR nicking enzyme amplification reaction

NO2 nitrobenzene

noCON DNA probe not containing p53CON

Os,bpy complex of osmium tetroxide with 2,2’-bipyridine

Os,L complex of osmium tetroxide with nitrogenous ligand

p53CON consensus binding site for the p53 protein

PAGE polyacrylamide gel electrophoresis

PCR polymerase chain reaction

PEX primer extension reaction

PGE pyrolytic graphite electrode

phen 1,10- phenanthroline

Pu purine

Py pyrimidine

SAE solid amalgam electrodes

SCS supercoil-selective DNA binding

SPR surface plasmon resonance

ssDNA single-stranded DNA

SWV square-wave voltammetry

TdT terminal deoxynucleotidyl transferase

TEMED N,N,N’,N’-tetramethyl ethylenediamine

69

TUNEL TdT-mediated dUTP-biotin nick end-labeling

V(D)J variable, diversity, and joining segments

wtp53 wild type p53

70

List of publications and conferences

1) Dual Redox Labeling as a Tool for Electrochemical Detection of p53 Protein-DNA

Interactions

Monika Hermanová, Petr Orság, Jana Balintová, Michal Hocek, Miroslav Fojta

2) Label-free voltammetric detection of products of terminal deoxynucleotidyl

transferase tailing reaction

Monika Hermanová, Pavlína Havranová, Anna Ondráčková, Swathi Senthil Kumar, Richard

Bowater and Miroslav Fojta

Electroanalysis (2018)

3) Butylacrylate-nucleobase Conjugates as Targets for Two-step Redox Labeling of

DNA with an Osmium Tetroxide Complex

Pavlína Havranová-Vidláková, Jan Špaček, Lada Vítová, Monika Hermanová, Jitka Dadová,

Veronika Raindlová, Michal Hocek, Miroslav Fojta and Luděk Havran

Electroanalysis (2018) 30, 371–377

Conferences:

Monika Hermanová, Jan Špaček, Petr Orság, Miroslav Fojta. Tail-labeled

oligonucleotide probes for a dual electrochemical magnetic immunoprecipitation assay of

DNA-protein binding. 11th

International Interdisciplinary Meeting on Bioanalysis, Brno,

Czech Republic, October 20-22 2014.

Monika Hermanová, Jan Špaček, Petr Orság, Miroslav Fojta. A new electrochemical

DNA binding competition assay using oligonucleotide substrates tail-labeled with two

different redox tags. CEITEC Annual Conference “Frontiers in Material and Life

Sciences”, Brno, Czech Republic, October 21-24 2014. Abstract Book, Page 131.

Monika Hermanová, Jan Špaček, Hana Pivoňková and Miroslav Fojta. Studies of

protein-DNA interactions using immunoprecipitation with DNA probes labelled with

71

electroactive groups. XXXV. Modern Electrochemical Methods, Jetřichovice, Czech

Republic, May 18-22 2015. Abstract Book, Pages 57-59.

Monika Hermanová, Jan Špaček, Miroslav Fojta. Electrochemical detection of DNA-

protein binding using dual redox tail-labeling of oligonucleotide probes. XXIII International

Symposium on Bioelectrochemistry and Bioenergetics of the Bioelectrochemical Society,

Malmö, Sweden, June 14-18 2015. Abstract Book, Page 215.

Monika Hermanová, Petr Orság, Miroslav Fojta. Elektrochemická detekce DNA-protein

interakcí využívající oligonukleotidy značené pomocí PEX reakce. XVI. Mezioborové setkání

mladých biologů, biochemiků a chemiků, Milovy, Czech Republic, May 10-13 2016.

Abstract Book, Page 65.

Monika Hermanová, Petr Orság, Miroslav Fojta. Electrochemical methods for detection of

protein-DNA interactions. XXV. Biochemický sjezd, Praha, Czech Republic, September 13-

16 2016. Abstract Book, Page 179.

Monika Hermanová, Petr Orság and Miroslav Fojta. Electrochemical methods for

detection of protein-DNA interactions. XXXVII. Modern Electrochemical Methods,

Jetřichovice, Czech Republic, May 15-19 2017. Abstract Book, Pages 58-61.

Monika Hermanová and Miroslav Fojta. Electrochemical analysis of products of terminal

deoxynucleotidyl transferase tailing reaction using intrinsic electroactivity of nucleobases. 69.

Zjazd chemikov, Vysoké tatry, Horný Smokovec, Slovakia, September 11-15 2017. Abstract

Book, Page 75.

Monika Hermanová, Pavlína Havranová, Anna Ondráčková and Miroslav Fojta.

Analysis of products of nontemplate enzymatic synthesis of DNA oligonucleotides using

voltammetric methods. XXXVIII. Modern Electrochemical Methods, Jetřichovice, Czech

Republic, May 21-25 2018. Abstract Book, Pages 76-79.

Monika Hermanová, Pavlína Havranová, Anna Ondráčková and Miroslav Fojta. Label-

free detection of products of terminal deoxynucleotidyl transferase tailing reaction. 17th

International Conference on Electroanalysis, Rhodes, Greece, June 3-7 2018. Abstract

Book, Page 156.

72

Appendices

73

Appendix 2

LABEL-FREE VOLTAMMETRIC DETECTION OF PRODUCTS OF TERMINAL

DEOXYNUCLEOTIDYL TRANSFERASE TAILING REACTION

Monika Hermanová1, Pavlína Havranová-Vidláková

1, Anna Ondráčková

1, Swathi Senthil

Kumar2, Richard Bowater

2, Miroslav Fojta

1,3,*

1 Institute of Biophysics, Czech Academy of Sciences, Kralovopolska 135, 612 65 Brno, Czech

Republic

2 School of Biological Sciences, University of East Anglia, Norwich, Norwich Research Park,

NR4 7TJ, United Kingdom

3 Central European Institute of Technology, Masaryk University, Kamenice 753/5, CZ-625 00

Brno, Czech Republic

*Corresponding author: [email protected]

74

Abstract

A label-free approach that takes advantage of intrinsic electrochemical activity of nucleobases

has been applied to study the products of terminal deoxynucleotidyl transferase (TdT) tailing

reaction. DNA homooligonucleotides A30, C30 and T30 were used as primers for the tailing

reaction to which a dNTP - or a mixture of dNTPs - and TdT were added to form the tails.

Electrochemical detection enabled study of the tailing reaction products created by various

combinations of primers and dNTPs, with pyrolytic graphite electrode (PGE) being suitable

for remarkably precise analysis of the length of tailing reaction products. Furthermore, the

hanging mercury drop electrode (HMDE) was able to reveal formation of various DNA

structures, such as DNA hairpins and G-quadruplexes, which influence the behavior of DNA

molecules at the negatively charged surface of HMDE. Thus, the described approach proves

to be an excellent tool for studying the TdT tailing reactions and for exploring how various

DNA structures affect both the tailing reactions and electrochemical behavior of DNA

oligonucleotides at electrode surfaces.

Key words: terminal deoxynucleotidyl transferase; oligonucleotide tailing; DNA

electrochemistry; label free; nucleobase; reduction; oxidation

75

1. Introduction

Terminal Deoxynucleotidyl Transferase (TdT) is a unique DNA polymerase as it catalyzes

random polymerization of deoxynucleotides at the 3'-OH ends of DNA in a template

independent manner [1,2]. The physiological role of TdT lies in random addition of

nucleotides to single-stranded DNA during V(D)J recombination. The ability of TdT to

randomly incorporate nucleotides increases antigen receptor diversity and, thus, TdT plays a

crucial role in the evolution and adaptation of the vertebrate immune system [3]. For its

catalytic activity, the enzyme requires a free 3'-OH group and a minimum of three nucleotide

residues to constitute the primer [4]. In vitro, TdT can incorporate all four natural

deoxyribonucleotides into single-stranded DNA. However, in vivo, a bias was observed

towards the incorporation of pyrimidine versus purine nucleotides [2], with the incorporation

efficiency of G being more than 4-fold higher than that of A [5,6]. TdT, like all DNA

polymerases, requires divalent metal ions for the catalysis. However, TdT is unique in its

ability to use a variety of divalent cations, such as Co2+

, Mn2+

, Zn2+

and Mg2+

[7]. Each metal

ion has different effects on the kinetics and mechanism of dNTP (deoxynucleoside

triphosphates; N stands for A, adenine, C, cytosine, G, guanine or T, thymine) utilization. For

example, Mg2+

facilitates the preferential utilization of dGTP and dATP whereas Co2+

increases the catalytic polymerization efficiency of the pyrimidines, dCTP and dTTP [8].

Nucleic acids are electrochemically active, which was first shown in the second half of the

1950s [9], and yield analytically useful signals at various electrodes [10,11]. Mercury-based

electrodes (such as hanging mercury drop electrode, HMDE) are traditionally used to measure

DNA signals related to reduction of its natural components (although it should be noted that

reduction of nucleobases at a graphite electrode has recently been reported on [12]). Here, we

use the HMDE as a well-established tool for DNA structure-sensitive measurements. In

mixed nucleotide sequences, adenine (A) and cytosine (C) yield the common cathodic CA

peak resulting from their reduction; guanine (G) gives the anodic GHMDE

peak, which results

from re-oxidation of a reduction product of guanine, 7,8-dihydroguanine, which is formed at

very negative potentials and, therefore, cannot be observed under usually applied conditions

due to an overlap of its signal with the cathodic background discharge at the mercury

electrode [13,14]. At carbon electrodes, A and G can be oxidized at positive potentials, giving

rise to AOX

and GOX

peaks, respectively. Analogues of natural purines, 7-deazaadenine (A*)

and 7-deazaguanine (G*), can also be studied using carbon electrodes as they produce

oxidation signals at positive potentials [15]. Moreover, adsorption/desorption processes or

76

reorientation of parts of the DNA chains occur at the negatively charged surface of mercury

electrodes, which can be observed as tensammetric (capacitive) signals [10]. Generally,

signals gained at mercury-containing electrodes are remarkably sensitive to changes in DNA

structure, which enables detection of different types of changes, such as formation of strand

breaks [16], DNA superhelicity-induced structural transitions [17], binding of DNA

intercalators [18] or formation of non-canonical DNA structures such as triplexes [19] and

quadruplexes [20,21].

Methods for studying the TdT tailing reaction have been based on polyacrylamide gel

electrophoresis (PAGE), usually along with radioactive labeling of DNA [22–27]. Here, for

the first time we apply an electrochemistry-based approach to study the TdT tailing reactions,

taking advantage of the intrinsic electroactivity of nucleobases. A set of DNA

homooligonucleotides were used as primers for the tailing reactions, to which a dNTP (or

a mixture of dNTPs) and TdT were added to form the tails. Taking account of the primer and

dNTP used in the particular reaction, an appropriate electrode was chosen for the tailing

reaction products analysis - HMDE or pyrolytic graphite electrode (PGE). We show that

electrochemical analysis enables a very precise study of the TdT tailing reaction products and,

furthermore, it can detect formation of DNA structures occurring in the TdT products.

2. Experimental section

2.1 Terminal Deoxynucleotidyl Transferase (TdT) reaction

As primers for the TdT reaction, single stranded oligonucleotides composed of one nucleotide

type, A30, C30 and T30 (Generi Biotech, Czech Rep.) were used. The reaction mixture

contained 5 M DNA primer (A30, C30 or T30), 5 M, 25 M, 50 M, 100 M or 250 M

dNTP (dGTP, dG*TP, dATP, dA*TP, dCTP or dTTP, Sigma-Aldrich, USA), 10 U Terminal

Deoxynucleotidyl Transferase (calf thymus, recombinant expressed in E. coli; New England

Biolabs, USA) and CoCl2 in a total volume of 20 l. Reaction mixtures were

incubated at 37 °C for 60 min. Products of the TdT reaction were purified using the

Nucleotide Removal Kit (Qiagen, the Netherlands) and dissolved in 30 l deionized water.

2.2 Denaturing PAGE

TdT reaction products were labeled by 32

P, dried out and dissolved in 5 l loading buffer

(80% formamide, 10 mM EDTA, 1 mg.ml-1

xylene cyanol FF, 1 mg.ml-1

bromphenol blue).

Aliquots (2.5 l) of the reaction mixtures were loaded into a 15% denaturing polyacrylamide

77

gel containing 1x TBE pH 8 and 7 M urea, which had been preheated at 25 W for 30 min, and

then electrophoresis continued at 25 W for 90 min. After drying, the gel was

autoradiographed and visualized using a Typhoon FLA 9000.

2.3 Electrochemical analysis

All electrochemical measurements were performed at room temperature in a three-electrode

setup (with the hanging mercury drop electrode, HMDE, or basal-plane pyrolytic graphite

electrode, PGE, as the working electrode, Ag/AgCl/3 M KCl as the reference electrode and

platinum wire as the auxiliary electrode) using an Autolab analyzer (Ecochemie, the

Netherlands) in connection with VA-Stand 663 (Metrohm, Switzerland). The measurements

were done in an adsorptive transfer stripping (AdTS) mode. DNA (10 g.ml-1

) was

accumulated at the electrode surface from 4 l aliquots (containing 0.2 M NaCl) for 60 s, then

the electrode was rinsed by deionized water and placed into the electrochemical cell

containing blank electrolyte. Cyclic voltammetry (CV) measurements at HMDE were

performed in 0.3 M ammonium formate with 0.05 M sodium phosphate, pH 6.9 as electrolyte;

CV settings were: initial potential 0.0 V, switching potential -1.85 V, final potential 0.0 V,

scan rate 1 V.s-1

, step potential 5 mV. Before each CV measurement the solution of the

background electrolyte was purged with argon. Square wave voltammetry (SWV)

measurements at PGE were performed in 0.2 M acetate buffer, pH 5.0; SWV settings were:

initial potential -1 V, final potential 1.6 V, frequency 200 Hz, amplitude 50 mV, scan rate 1

V.s-1

. Before each SWV measurement the PGE was pretreated by applying a potential of 1.8

V for 30 s in the background electrolyte, and its surface was renewed by peeling off the

graphite top layer using sticky tape.

3. Results and discussion

For the tailing reaction, three types of homooligonucleotides – A30, C30 and T30 – were used as

primers to which various dNTPs (dATP, dCTP, dGTP, dTTP, 7-deazaadenine - dA*TP and 7-

deazaguanine - dG*TP or their combinations) were added to form the tail. G30 was not used as

a primer since DNA oligonucleotides containing longer stretches of guanines tend to form G-

quadruplexes (G4) [21,28] and such structure would not be a suitable substrate for TdT

because of its inability to access the 3'-OH end folded within the G4 structure. Figure 1 shows

cyclic voltammograms at HMDE and square wave voltammograms at PGE gained for the A30,

C30 and T30 primers compared to blank electrolyte. For the A30 primer, the A reduction peak at

HMDE (AHMDE

) and A oxidation peak at PGE (AOX

) can be observed. C30 yields only a

78

reduction peak at HMDE. No faradaic signals can be observed for T30 at HMDE or PGE

under these conditions. However, with HMDE, a tensammetric peak (TTENS

) resulting from

the desorption and reorientation processes of T30 and taking place at negative potentials

(around -1.5 V) at mercury electrode surface is evident. This behavior of oligo(dT) has been

described previously [29]. Both cyclic and square wave voltammograms of the individual

primers show that when measuring the particular homooligonucleotide primer, signals of no

other bases are visible, which means that after choosing an appropriate combination of primer

and dNTP, the primers can be used for monitoring the TdT tailing reaction. The TdT tailing

reaction was performed in five different dNTP:primer molar ratios: 1:1, 5:1, 10:1, 20:1 and

50:1; a “no enzyme” control was performed in order to compare signals gained for the tailing

reaction products to those of the sole primer and to exclude interference of dNTP residues

possibly present in the tailing product solutions.

Figure 1. Cyclic voltammograms at HMDE (A) and square wave voltammograms at PGE (B)

of the A30, C30 and T30 primers. Signals yielded by individual primers are labelled.

79

Figure 2. Examples of sections of cyclic (obtained at HMDE) and square wave (obtained at

PGE) voltammograms gained for various combinations of primers and dNTPs: (A) A30+G

HMDE; (B) A30+G PGE; (C) A30+G* PGE; (D) A30+A* PGE; (E) C30+A HMDE; (F) C30+A

PGE; (G) A30+C HMDE; (H) T30+C HMDE. Reaction conditions and significant conclusions

are discussed in the text.

Guanine (G) and 7-deazaguanine (G*) were used for the tailing reaction with all three primers

- A30, C30 and T30 (Figure 2). Regardless of which primer was used, the signals observed

corresponded with composition of the reaction mixture i.e., for dGTP peak GHMDE

and peak

GOX

were observed, whereas for dG*TP peak G*OX

appeared. The A30+G combination

revealed a significant difference between the results gained at HMDE and PGE for the same

80

reaction products (Figures 2A, B and 3A). With PGE the height of the GOX

peak continuously

increased with the increasing dNTP:primer ratio. In contrast, peak GHMDE

showed a

decreasing tendency from a 5:1 to 50:1 ratio after the initial growth between 1:1 and 5:1 ratio.

This behavior at HMDE is in contrast to what was observed at PGE. An analogous effect,

decrease of peak GHMDE

with increasing length of Gn stretches, has been previously observed

and ascribed to the formation of intermolecular parallel G-quadruplexes (G4) [21]. Since the

G-rich sequences are generally known to adopt the structure of G4 and this ability is more

pronounced in longer Gn stretches, it can be assumed that the G-tails formed G4 at

dNTP:primer molar ratios of 5:1 and higher. This assumption is further supported by the fact

that K+ ions, which stabilize formation of G4, are present in the TdT reaction buffer, in which

the tailing reaction takes place.

The denaturing PAGE autoradiogram showed that the length of the tailing reaction products

did not grow significantly at dGTP:primer ratios higher than 5:1. Moreover, for the ratios 5:1

and higher, heavier species were formed (detected at starts of the gel, see Figure 3A), whose

mass was too big to enter the gel; these species can be expected to be formed by the

intermolecular G4. This suggested that DNA obtained from the G-tailing reactions occurred in

two states: G4, which were resistant to the denaturing conditions during the PAGE, and

unstructured oligonucleotides that create a ladder typical for TdT tailing reaction products. In

principle, formation of G4 structures influences the intensity of the measured peak GHMDE

in

two ways. Firstly, when the dNTP:primer ratio is sufficient to produce G-tails that are capable

of stable G4 formation, the tailing reaction is stopped when the G4 is formed because of the

inaccessibility of the 3'-OH end to TdT. Secondly, the G4 formation changes behavior of the

oligonucleotides at the mercury electrode surface, which is related to the fact that DNA

structures existing in solution can be preserved after adsorption at the electrode surface in

adsorptive transfer techniques [10]. Moreover, mercury-containing electrodes are sensitive to

changes in DNA structure. It has been well established that the accessibility of nucleobase

residues for electrode reactions is strongly dependent on the DNA structure, particularly at

mercury electrodes. It can be assumed that this feature is responsible for the decrease of the

signals at the HMDE, together with the fact that electrostatic repulsion of the rigid quadruplex

structures from the negatively charged surface of the HMDE may hinder reduction of guanine

residues involved [21]. On the other hand, the oxidation of guanine at the PGE takes place at

the positively charged surface of the electrode to which the oligonucleotide molecules are

attracted. This facilitates the oxidation process even in the quadruplex structure. Along with

bigger accessibility of the primary oxidation site of guanine to the electrode surface, the

81

signals gained for oxidation of guanine at the PGE depend merely on the length of the

oligonucleotide molecules and do not reflect any structural changes occurring in the

molecules. When comparing the PAGE autoradiogram and the GOX

peak heights at the PGE,

very similar trends were observed, with a steep increase between the 1:1 and 5:1 ratio

followed by a slower increase from 5:1 to 50:1 ratio. PAGE and electrochemical analysis with

the PGE thus provided comparable results as to the length of the tailing reaction products,

thus making the latter a suitable alternative method for studying the process of the TdT tailing

reactions. On the other hand, electrochemical measurements in the negative potential region

with the HMDE provide extra information about DNA structural transitions.

82

Figure 3. Heights of the respective peaks gained at HMDE or PGE and denaturing PAGE

autoradiograms obtained for combinations of G and G* with all three primers: (A) A30+G; (B)

A30+C; (C) T30+G; (D) A30+G*; (E) C30+G*; (F) T30+G*. Reaction conditions and significant

conclusions are discussed in the text.

In the case of T30+G (Figure 3C), the denaturing PAGE revealed the same pattern of the

length of the tailing reaction products as A30+G, indicating formation of G4 at ratios higher

than 5:1, and a very similar trend as for A30+G was also observed using voltammetric analysis

at PGE. Nevertheless, voltammetric responses of the T30+G tailing products at HMDE could

not be compared to those of A30+G because behavior of the former at HMDE is

predominantly affected by the T30 oligonucleotide due to strong interactions of

homopyrimidine stretches with the negatively charged HMDE surface [29], which can further

affect adsorption and redox processes of the G-tails, especially if G4 are formed. These

phenomena, whose detailed description is out of scope of this report, are intensively studied in

our laboratory (S. Hasoň, H. Pivoňková et al., manuscripts in preparation).

Unlike A30+G and T30+G, the C30+G tailing reaction products (Figure 3B) did not form G4,

which is evident from the PAGE autoradiogram showing that the tail length does not stop

growing after a particular ratio and the longest products are obtained for the highest ratio

(50:1). Voltammetric analysis at PGE again reveals a good correlation with the PAGE

analysis. However, at HMDE, the signal grows only until the 20:1 ratio, then for the 50:1 ratio

the signal drops significantly. This can be explained by formation of another DNA structure,

such as a DNA hairpin (or, alternatively, duplex dimer of self-complementary GnCm

oligonucleotides). Bases C and G are complementary and tend to form duplex DNA when

possible, therefore when the G-tail reaches sufficient length for the hairpin formation (in this

case around 30 added nucleotides) it can bend over and form a hairpin with the

complementary C30 primer. The formation of such alternative structures can make signals at

HMDE decrease. It can also explain why only one band corresponding to two or three lengths

of the tailing products occurs in the PAGE autoradiogram, suggesting that the tailing reaction

stops when the DNA hairpin is formed.

For comparison with the behavior of the natural guanine, 7-deazaguanine (G*) was used,

which is an analogue of natural guanine in which the N7 atom is replaced by CH group. This

feature implicates that while the Watson-Crick base pairing (with C) of the 7-deazaguanine

83

remains unaffected, the Hoogsteen base pairs cannot be formed due to the involvement of the

N7 atom in the formation of this type of base pair. Thus, DNA molecules containing G* are

incapable of forming multi-stranded DNA structures, such as triplexes or quadruplexes. This

fact is utilized, for example, in PCR amplification of G-rich sequences where the quadruplex

formation could affect the amplification process. For analysis of the A30+G* tailing reaction

products (Figure 2C and 3D), only oxidation at PGE was used since, with HMDE, G* does

not yield any peak analogous to the GHMDE

peak, in agreement with absence of the

corresponding redox site in G* [30]. Both voltammetric and PAGE analysis showed that,

unlike in the case of A30+G tailing products, the length of the tailing products grew gradually

with the rising dG*TP:primer ratio, with the average tail length much higher than in the case

of dGTP, especially for the 20:1 and 50:1 ratios and the G*OX

signal for the 50:1 ratio being

~10-fold higher than that for the 1:1 ratio. This was in agreement with the fact that, unlike

natural guanine, G* is not capable of forming G4 and, therefore, the TdT tailing reaction was

not inhibited in contrast to what was observed in experiments with G (see above). G* is thus

suitable for studying how TdT incorporates an analogue of G into DNA as it does not exhibit

limitations related to formation of G4 structures described above for G.

Practically the same results as for the A30+G* were obtained for the T30+G* tailing products

(Figure 3F) with a similar pattern of the tailing products lengths visible on the PAGE

autoradiogram and intensities of the G*OX

signal gained at PGE. On the other hand, the

pattern observed for C30+G* (Figure 3E) resembled more that of the C30+G tailing products

than those of A30+G* and T30+G*, which is in accordance with the fact that the ability of G*

to form Watson-Crick pairs with C is not violated, which allows hairpin formation after a

certain length of the tail is reached.

84

Figure 4. Heights of the respective peaks gained at HMDE or PGE and denaturing PAGE

autoradiograms obtained for combinations of A and A* with all three primers: (A) T30+A; (B)

C30+A; (C) C30+A*; (D) A30+A*. Reaction conditions and significant conclusions are

discussed in the text.

Besides guanine and 7-deazaguanine, adenine and 7-deazaadenine (A*) were also used for the

tailing reactions, together with primers T30, C30 and A30 (the latter only for A*). For the T30+A

(Figure 4A), PAGE analysis revealed that for the 1:1 ratio an average of two nucleotides were

added. However, starting at the 5:1 ratio, the reaction stopped after 5 or 6 added nucleotides.

Therefore, all the reaction products for the ratios 5:1, 10:1, 20:1 and 50:1 had tails consisting

of 5 or 6 nucleotides. This behavior might be caused by a formation of a DNA hairpin (or

dimer) since A and T are complementary bases and, like C and G, they are prone to duplex

creation. When the A tail reaches sufficient length for the hairpin formation, the DNA bends

and a DNA hairpin with a 5' overhang is created. This means that the 3'-OH end becomes

poorly accessible for the TdT and the enzyme cannot continue in the tailing reaction – at least

in certain population of the tailing products. Electrochemical results obtained for the T30+A

again displayed significant difference between the data obtained at HMDE and PGE

(similarly to A30+G and C30+G). In the cyclic voltammogram gained at HMDE, the

85

tensammetric peak mentioned above caused by the T30 primer can be observed. The potential

of the TTENS

peak coincides with that of AHMDE

peak. Thus, these two peaks cannot be

distinguished and there is a non-zero value for the primer (no TdT control). Then, for the 1:1

and 5:1 ratios the height of the collective peak at -1.52 V increased in comparison to the

signal gained for the sole primer, which agreed with the growing length of the A tails

observed at the PAGE autoradiogram. For the higher ratios, where the DNA hairpin is

expected to be formed, the signal sharply decreased, which was also observed for C30+G and

is in agreement with the fact that HMDE is sensitive to structural changes occurring in DNA.

On the other hand, the signals gained at PGE generally exhibit a poor sensitivity to variations

in DNA structure; therefore, the height of the AOX

peak displays a trend very similar to that

visible at the PAGE autoradiogram.

The C30+A combination generates a specific signal for the bases forming the primer and the

tail (Figure 2E,F and 4B). This occurs because in DNAs with a random distribution of bases,

C and A yield a common CA peak at HMDE since their reduction potentials are in close

proximity. However, in DNA composed of C and A homooligonucleotide stretches, two

separated peaks can be observed under certain conditions [31–33]. In this case, it was not

possible to see the separate C and AHMDE

peaks, only a hint of the AHMDE

peak appeared at a

potential more negative than that of the C peak for the product with the longest A tail (i.e. for

the 50:1 ratio). The height of the C peak decreased with the increasing dATP:primer ratio,

reflecting the decreasing proportion of cytosines in the tailing products. At PGE, the intensity

of the AOX

peak increased with higher ratios of dATP:primer, but it was not possible to

compare the trend of this signal to the pattern gained at PAGE since, in this case, the PAGE

analysis did not show distinct bands of the tailing products. Most probably this was caused by

a wider distribution of the lengths of tails, leading to smaller amounts of DNA of each

specific length, meaning that the bands corresponding to such lengths could not be visualized.

Nevertheless, the intensity of the band of the unextended C30 primer decreased with the higher

ratios of dATP:primer, which indicates that the primer has been consumed for the tailing

reaction. The same effect can be observed at the PAGE autoradiogram of the C30+A* tailing

products (Figure 4C), again making the electrochemical analysis at PGE more accurate than

the PAGE analysis. In both cases, the general trends of these signals are practically the same

(unlike in the case of A30+G), indicating that no unusual structure is formed in the A tail.

Thus, we can see that in some instances electrochemical analysis at PGE responds to the

increase of mean length of the tailing products more clearly than the PAGE analysis.

86

Results gained for the A30+A* tailing products (Figure 2D and 4D) show that the trend of the

growing tail length visible at both PAGE autoradiogram and electrochemical analysis at PGE

is different from that of C30+A* tailing products, which probably reflects different preferences

of the TdT for different primers and dNTP substrates. However, again there is a striking

agreement between the results of PAGE and electrochemistry with PGE for the given primer-

dNTP pair.

Figure 5. Heights of the respective peaks gained at HMDE and denaturing PAGE

autoradiograms obtained for combinations of C and T with all three primers: (A) T30+C; (B)

A30+C; (C) A30+T; (D) C30+T. Reaction conditions and significant conclusions are discussed

in the text.

PAGE analysis of tailing products having tails consisting of cytosines (T30+C – Figure 2H and

5A and A30+C – Figure 2G and 5B) showed that the length of the products grows with the

increasing ratio of dCTP:primer, but even at the highest ratios the products do not appear to

be as long as for other bases. Since the incorporated cytosines cannot be analyzed using PGE

under the used conditions, cyclic voltammetry at HMDE was used for their detection. For

T30+C, the same tensammetric peak caused by the T30 primer (TTENS

) as in the case of T30+A

87

was observed. Although TTENS

appeared at a potential close to that of the C peak, it was

possible to distinguish these two peaks as the potential at which it occurred was about 400

mV more negative than that of the C peak. The T30 peak was clearly discernible only in the

“no enzyme” control and the 1:1 ratio, with a slight hint at the 5:1 ratio; for the higher ratios

the peak diminishes, indicating that with the increasing ratio of dCTP:primer, the behavior of

the C-tail at the electrode surface prevails. The height of the C peak then grows gradually with

the increasing ratio. In the case of A30+C, similar situation occurred as for the C30+A, with C

and A being reduced at HMDE at close potentials. However, here it was possible to observe a

separate C peak for the higher ratios of dCTP:primer. With an increasing number of cytosines

forming the tail, it was observed that the C peak appeared at a less negative potential than the

AHMDE

peak. For the 50:1 ratio, the peak was already well developed. The peak heights gained

for the CA peak showed no obvious trend and remained similar for the primer and all ratios of

dCTP:primer. Nevertheless, an evident tendency to be higher with the increasing ratio was

seen when the C peak was analyzed separately. Moreover, this tendency matched the results

gained from the PAGE analysis, with a gradual increase from the 1:1 to 10:1 ratio and further

increase between 20:1 and 50:1 ratio.

Primers A30 and C30 were also used for tailing reaction with dTTP. The PAGE analysis of the

A30+T tailing products (Figure 5C) showed, that unlike for T30+A, the tailing reaction was not

stopped after a few added nucleotides and the tailing products gradually increased in length

up to the 50:1 ratio. It remains unclear why the behavior of TdT during the A30+T tailing

reaction differs from that of the T30+A combination. When analyzed at HMDE, the height of

the A peak displayed a decreasing tendency, which reflects a decreasing amount of the A30

parts of the tailing products that were adsorbed at the electrode surface due to a growing

portion of the T-tails. The PAGE analysis of the C30+T (Figure 5D) revealed the same

problem as for the C30+A and C30+A* combinations – no distinct tailing products at the

autoradiogram. However, the diminishing intensity of the C30 band indicates that the tailing

products were created but their length cannot be assessed. The voltammetric analysis of

C30+T at HMDE confirmed that the cytosines adsorb very strongly at the mercury surface

[29], which is evident from the fact that the C peak height remains unchanged for all ratios of

dTTP:primer.

Figure 6 shows the relative intensities of the peaks specific for the particular bases obtained

using square wave voltammetry at PGE or cyclic voltammetry at HMDE. Values of the peak

heights were normalized to the 1:1 dNTP:primer ratio, which enabled comparison of the

88

relative signals gained for the individual combinations. The highest relative signal intensities

were gained for four different combinations of primers and dNTPs – C30+G, C30+A, C30+A*

and T30+C – indicating that the mutual combination of the primer and used dNTP and their

ability (or ability of the tail alone) to form various structures dictate the resulting outcomes of

the tailing reactions. High relative signals were also obtained for primer+dNTP combinations

in which equimolar mixtures of two different dNTPs were used for the tailing reactions,

suggesting that mixing two dNTPs has an impact on formation of possible DNA structures (in

this case G4). Comparing the normalized GHMDE

peak height gained for A30+G+C and

A30+G+T, we see that unlike dCTP, adding dTTP to the reaction with dGTP did not prevent

the guanines in the tail from creating the G4 structure, which is evident from the decrease of

the peak height for the 50:1 ratio for the A30+G+T products. This difference is probably

caused by the ability of G and C to form duplex DNA, which reduces G stretches present in

the G+C tails that would otherwise tend to form G4, and by the preference of the TdT reaction

for incorporation of G in competition with T [5,6]. For more details on A30+G+C, A30+G+T

and A30+G*+C results, see SI.

Figure 6. Normalized intensities of the respective peaks gained for all combinations of

primers and dNTPs used for the TdT tailing reaction, including combinations with two mixed

dNTPs (A30+G+C, A30+G+T and A30+G*+C).

4. Conclusions

We used intrinsic electroactivity of nucleobases at HMDE and PGE to study products of the

TdT tailing reactions and show that we can monitor how the tailing reaction works for various

89

combinations of primers and dNTPs. This approach enabled study of the tailing reaction and

also allowed observation of the formation of DNA structures that may affect the tailing

reactions and electrochemical behavior of the respective oligonucleotides. This was

particularly evident in the case of HMDE since, for all tailing products capable of forming

unusual DNA structures, we observed a decrease in the corresponding signal, which reflected

the structural changes and their impact on the behavior of these molecules at the negatively

charged surface of HMDE. On the other hand, results obtained at PGE corresponded very

precisely to those obtained using denaturing PAGE, which implies that PGE can be used for

monitoring the lengths of the TdT tailing reaction products without dramatic interfering

effects of secondary DNA structure. Thus, combination of measurements at both types of

electrodes can provide complete (and complementary) information. In some cases, the

electrochemical analysis at PGE was even more accurate than the denaturing PAGE. Thus,

this approach proves to be an excellent tool for studying the TdT tailing reactions and also for

exploring electrochemical behavior of the DNA oligonucleotides at electrode surfaces.

5. Acknowledgements

This work was supported by the Czech Science Foundation (grant P206/12/G151), by the

ASCR (RVO 68081707), by the Ministry of Education, Youth and Sports of the Czech

Republic under project CEITEC 2020 (LQ1601), by the SYMBIT project reg. no.

CZ.02.1.01/0.0/0.0/15_003/0000477 financed from the ERDF, and from the European

Union’s Horizon 2020 research and innovation programme (project No 692068 BISON).

6. References

[1] K.-I. Kato, J. M. Goncalves, G. E. Houts, F. J. Bollum, J. Biol. Chem. 1967, 242,

2780–2789.

[2] E. A. Motea, A. J. Berdis, Biochim. Biophys. Acta - Proteins Proteomics 2010, 1804,

1151–1166.

[3] D. Baltimore, Nature 1974, 248, 409–411.

[4] F. J. Bollum, Enzymes 1974, 10, 145–171.

[5] B. Yang, K. N. Gathy, M. S. Coleman, J. Biol. Chem. 1994, 269, 11859–11868.

[6] M. J. Modak, Biochemistry 1978, 17, 3116–3120.

[7] M. R. Deibel, M. S. Coleman, J. Biol. Chem. 1980, 255, 4206–12.

[8] M. S. Chang, F. J. Bollum, J. Biol. Chem. 1990, 265, 17436–17440.

90

[9] E. Paleček, Naturwissenchaften 1958, 45, 186–187.

[10] E. Paleček, M. Bartošík, Chem. Rev. 2012, 112, 3427–3481.

[11] M. Fojta, F. Jelen, L. Havran, E. Paleček, Curr. Anal. Chem. 2008, 4, 250–262.

[12] J. Špaček, A. Daňhel, S. Hasoň, M. Fojta, Electrochem. Commun. 2017, 82, 34–38.

[13] L. Trnková, M. Studničková, E. Paleček, J. Electroanal. Chem. 1980, 116, 643–658.

[14] M. Studničková, L. Trnková, J. Zetěk, Z. Glatz, Bioelectrochem. Bioenerg. 1989, 275,

83–86.

[15] H. Pivoňková, P. Horáková, M. Fojtová, M. Fojta, Anal. Chem. 2010, 82, 6807–6813.

[16] M. Fojta, E. Palecek, Anal. Chim. Acta 1997, 342, 1–12.

[17] M. Fojta, R. P. Bowater, V. Staňková, L. Havran, D. M. J. Lilley, E. Paleček,

Biochemistry 1998, 37, 4853–4862.

[18] M. Fojta, L. Havran, J. Fulnečková, T. Kubičárová, Electroanalysis 2000, 12, 926–934.

[19] X. Cai, G. Rivas, H. Shirashi, P. Farias, J. Wang, M. Tomschik, F. Jelen, E. Paleček,

Anal. Chim. Acta 1997, 344, 65–76.

[20] A. M. Chiorcea-Paquim, A. M. Oliveira-Brett, Electrochim. Acta 2014, 126, 162–170.

[21] P. Vidláková, H. Pivoňková, I. Kejnovská, L. Trnková, M. Vorlíčková, M. Fojta, L.

Havran, Anal. Bioanal. Chem. 2015, 407, 5817–5826.

[22] R. Roychoudhury, E. Jay, R. Wu, Nucleic Acids Res. 1976, 3, 863–77.

[23] G. Deng, R. Wu, Nucleic Acids Res. 1981, 9, 4173–4188.

[24] A. M. Michelson, S. H. Orkin, J. Biol. Chem. 1982, 257, 14773–14782.

[25] J. B. Boulé, F. Rougeon, C. Papanicolaou, J. Biol. Chem. 2001, 276, 31388–31393.

[26] J. Gouge, S. Rosario, F. Romain, P. Beguin, M. Delarue, J. Mol. Biol. 2013, 425,

4334–4352.

[27] J. Loc’h, S. Rosario, M. Delarue, Structure 2016, 24, 1452–1463.

[28] E. P. Wright, J. L. Huppert, Z. A. E. Waller, Nucleic Acids Res. 2017, 45, 2951–2959.

[29] S. Hasoň, V. Vetterl, M. Fojta, Electrochim. Acta 2008, 53, 2818–2824.

[30] Z. Dudová, J. Špaček, M. Tomaško, L. Havran, H. Pivoňková, M. Fojta, Monatsh.

Chem. 2016, 147, 3–11.

[31] L. Trnková, F. Jelen, I. Postbieglová, Electroanalysis 2003, 15, 1529–1535.

[32] L. Trnková, F. Jelen, I. Postbieglová, Electroanalysis 2006, 18, 662–669.

[33] R. Mikelová, L. Trnková, F. Jelen, V. Adam, R. Kizek, Electroanalysis 2007, 19, 348–

355.

91

Appendix 3

92

93

94

95

96

97