Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19

Contents lists available at SciVerse ScienceDirect

Journal of Photochemistry and Photobiology B: Biology


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

Short Review

Drug–DNA interactions and their study by UV–Visible, fluorescence


spectroscopies and cyclic voltametry
Muhammad Sirajuddin 1, Saqib Ali ⇑, Amin Badshah 2
Department of Chemistry, Quaid-i-Azam University, Islamabad 45320, Pakistan

a r t i c l e i n f o a b s t r a c t

Article history: The present paper review the drug–DNA interactions, their types and applications of experimental tech-
Received 7 February 2013 niques used to study interactions between DNA and small ligand molecules that are potentially of phar-
Received in revised form 26 March 2013 maceutical interest. DNA has been known to be the cellular target for many cytotoxic anticancer agents
Accepted 28 March 2013
for several decades. Understanding how drug molecules interact with DNA has become an active research
Available online 6 April 2013
area at the interface between chemistry, molecular biology and medicine. In this review article, we
attempt to bring together topics that cover the breadth of this large area of research. The interaction
Keywords:
of drugs with DNA is a significant feature in pharmacology and plays a vital role in the determination
Drug–DNA interaction
UV–Visible spectroscopy
of the mechanisms of drug action and designing of more efficient and specifically targeted drugs with les-
Fluorescence spectroscopy ser side effects. Several instrumental techniques are used to study such interactions. In the present
Cyclic voltammetry review, we will discuss UV–Visible spectroscopy, fluorescence spectroscopy and cyclic voltammetry.
The applications of spectroscopic techniques are reviewed and we have discussed the type of information
(qualitative or quantitative) that can be obtained from the use of each technique. Not only have novel
techniques been applied to study drug–DNA interactions but such interactions may also be the basis
for the development of new assays. The interaction between DNA and drugs can cause chemical and con-
formational modifications and, thus, variation of the electrochemical properties of nucleobases.
Ó 2013 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Drug–DNA interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3. Types of Drug–DNA interaction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
4. Modes of Drug–DNA binding. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
4.1. Covalent mode of binding. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
4.1.1. Alkylating agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
4.2. Non-covalent mode of binding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
4.2.1. Intercalation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
4.2.2. Groove binding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4.2.3. External binding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5. Techniques used to study Drug–DNA interactions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5.1. UV–Visible absorption spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
5.2. Fluorescence emission spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
5.3. Cyclic voltammetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
6. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
7. Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

⇑ Corresponding author. Tel.: +92 51 90642130; fax: +92 51 90642241.


E-mail addresses: m.siraj09@yahoo.com (M. Sirajuddin), drsa54@yahoo.com (S. Ali), aminbadshah@yahoo.com (A. Badshah).
1
Tel.: +92 51 90642208; fax: +92 51 90642241.
2
Tel.: +92 51 90642131; fax: +92 51 90642241.

1011-1344/$ - see front matter Ó 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jphotobiol.2013.03.013
2 M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19

1. Introduction via alkylation or inter- and intrastrand crosslinking [8]. The cova-
lent mode of binding of drug–DNA is irreversible and invariably
DNA is present in body in the form of a double helix (Fig. 1A), causes the complete inhibition of DNA processes and subsequent
where each strand is composed of a combination of four nucleo- cell death (if the reaction is not directly chemically reversible). A
tides (a nucleotide is a nucleoside with one or more phosphate major advantage of covalent binders is the high binding strength.
groups covalently attached to the 30 - and/or 50 -hydroxyl group(s)), Moreover, covalent bulky adducts can cause DNA backbone distor-
adenine (A), thymine (T), guanine (G) and cytosine (C) (Fig. 1B). tion, which in turn can affect both transcription and replication,
Within a strand these nucleotides are connected through phospho- such as by disrupting protein complex recruitment [9].
diester (deoxyribose sugars joined at both the 30 -hydroxyl and Cis-platin [cis-dichlorodiammineplatinum(II)] is a famous cova-
50 -hydroxyl groups to phosphate groups in ester links is called lent binder used as an anticancer drug, and makes an intra/inter-
phosphodiester linkage) linkages (Fig. 1C). The two strands are strand cross-link through the chloro groups with the nitrogens
held together primarily through Watson Crick hydrogen bonds on the DNA bases (Scheme 2).
where A forms two hydrogen bonds with T and C forms three The covalent attachment of organic intercalators to transition
hydrogen bonds with G (Fig. 1D) [1–3]. The region where the metal complexes, producing metallointercalators, can lead to novel
two strands are close to each other (deep–narrow) is called minor DNA interaction that affect biological activity. Metal complexes hav-
grove while the region where they are away from each other ing r-bonded aromatic side arms can acts as dual-function com-
(shallow–wide) is called major groove. plexes: they bind to DNA both by metal coordination and through
intercalation of the attached aromatic ligand. These aromatic side
2. Drug–DNA interactions arms introduce new mode of DNA binding, involving mutual inter-
actions of functional groups held in close proximity [10].
DNA is the pharmacological target of many of the drugs that are The covalent binders are also called alkylating agents due to ad-
currently in clinical use or in advanced clinical trials. Targeting duct formation because they are used in cancer treatment to attach
DNA to regulate cell functions by modulating transcription (gene an alkyl group (CnH2n+1) to DNA [11].
expression and protein synthesis) or by interfering with replication
(a major step in cell growth and division) seems logical, intuitively 4.1.1. Alkylating agents
appealing and conceptually straightforward. Small ligand mole- DNA alkylating agents have been used as anticancer agents by pro-
cules bind to DNA and artificially alter and/or inhibit the function- ducing significant DNA damage to kill cancer cell [12]. Alkylating
ing of DNA. These small ligand molecules act as drug when agents involve reactions with guanine in DNA. However, some alkylat-
alteration or inhibition of DNA function is required to cure or con- ing drugs exert their strongest effects through alkylation of other posi-
trol a disease [4]. tions and other types of cross-links. These drugs add methyl or other
The study of interaction of drug with DNA is very exciting and alkyl groups onto molecules where they do not belong. This in turn
significant not only in understanding the mechanism of interac- inhibits their correct utilization by base pairing and causes a miscod-
tion, but also for the design of new drugs. However, mechanism ing of DNA. Alkylating agents can interact via three mechanisms.
of interactions between drug molecules and DNA is still relatively In the first mechanism, an alkylating agent attaches alkyl groups
little known. It is necessary to introduce more simple methods for to DNA bases. This alteration results in the DNA being fragmented by
investigating the mechanism of interaction. By understanding the repair enzymes in their attempts to replace the alkylated bases.
mechanism of interaction, designing of new DNA-targeted drugs A second mechanism by which alkylating agents cause DNA
and the screening of these in vitro will be possible [5]. damage is the formation of cross-links, bonds between atoms in
the DNA. In this process, two bases are linked together by an alkyl-
ating agent that has two DNA binding sites. The cross-linking of the
3. Types of Drug–DNA interaction two strands of DNA produced by the bifunctional alkylating agents
prevents the use of that DNA as a template for further DNA and
There are primarily three different ways by which anticancer RNA synthesis leading to inhibition of replication and transcription
drugs interact with DNA (Scheme 1), (i) through control of tran- and, then, to cell death. A large number of chemical compounds are
scription factors and polymerases. Here, the drugs interact with alkylating agents under physiological conditions, and a variety of
proteins that bind to DNA, (ii) through RNA binding to DNA double such compounds have exhibited antitumor activity.
helix to form nucleic acid triple helix structures or RNA hybridiza- The third mechanism of action of alkylating agents causes the
tion (sequence specific binding) to exposed DNA single strand mispairing of the nucleotides leading to mutations. The alkylating
forming DNA–RNA hybrids that may interfere with transcriptional agents, by chemical interactions, form covalent links with DNA.
activity and (iii) through binding of small aromatic ligand mole- This causes ‘‘mistakes’’ in the DNA that may result in mispairing,
cules to DNA double helical structures. The binding of small mole- substitutions, or excision. The cellular response of these mistakes
cules to DNA involves electrostatic interaction, intercalation may inhibit DNA synthesis and proliferation or they may result
between base pairs and minor and major DNA grooves binding in apoptosis (process of programmed cell death that may occur
interaction [6]. in multinuclear organisms) [11].
Alkylating drugs are the oldest class of anticancer drugs still
4. Modes of Drug–DNA binding commonly used; they play an important role in the treatment of
several types of cancers. Most alkylating drugs are monofunctional
There are two modes of drug–DNA binding, covalent and non- methylating agents (e.g., temozolomide [TMZ], N-methyl-N0 -nitro-
covalent. Some examples of DNA-interacting agents are given in N-nitrosoguanidine [MNNG], and dacarbazine), bifunctional alkyl-
Table 1[7]. ating agents such as nitrogen mustards (e.g., chlorambucil and
cyclophosphamide), or chloroethylating agents (e.g., nimustine
4.1. Covalent mode of binding [ACNU], carmustine [BCNU], lomustine [CCNU], and fotemustine)
[13]. The alkylation of N3 of guanine by (+)-CC1065 (antitumor
Many anticancer drugs in clinical use function by interacting antibiotic) is shown in Scheme 3 [14]. (+)-CC1065 induces both
with DNA, not only those that bind to DNA through non-covalent DNA bending toward the minor grooves and widing equivalent to
interactions but also those that form covalent adducts such as about one base pair per covalent alkylation event. (Scheme 4)
M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19 3

Fig. 1. (A) Structure of DNA molecule, (B) bases (adenine, guanine, thymine and cytosine) of DNA. (C) Watson–Crick pairing between purine and pyrimidine bases in
complementary DNA strands. (D) Watson Crick base pairing, A–T and G–C base pairing. Arrows indicates potential nitrogen and oxygen coordination sites.

4.2. Non-covalent mode of binding on convoluted molecular and biochemical pathways is not well
characterized, the most important effects include DNA conforma-
Non-covalent DNA interacting agents, DNA-groove binders and tional and related structural perturbations, interference with nor-
DNA intercalators, are generally considered less cytotoxic than mal DNA protein interactions, such as topoisomerases, as well as
agents producing covalent DNA adducts and other DNA damage. effects on mitochondrial DNA and function. The non-covalent bind-
Although the impact of non-covalent compound–DNA interactions ing mode is reversible and is typically preferred over covalent
4 M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19

Scheme 1. Summary of mechanism of action of anticancer drugs.

adduct formation keeping the drug metabolism and toxic side effects in tors are used in chemotherapeutic treatment to inhibit DNA repli-
mind. Non-covalent DNA interacting agents can change DNA confor- cation in rapidly growing cancer cells. DNA-intercalator complex is
mation, change DNA torsional tension, interrupt protein–DNA interac- stabilized by p–p stacking interaction and is thus less sensitive to
tion, and potentially lead to DNA strand breaks. All of these events can ionic strength relative to the two other binding modes (groove
have substantial effects on gene expression [8]. binding and external binding). Structural changes are induced in
The non-covalent mode of drug–DNA binding is further classi- DNA by intercalators. Intercalation stabilizes, lengthens, stiffens,
fied into three types, intercalation, groove binding and external and unwinds the DNA double helix [22]. In order for an intercalator
binding (on the outside of the helix) [15]. to fit between base pairs, the DNA must dynamically open a space
between its base pairs by unwinding. The degree of unwinding var-
4.2.1. Intercalation ies depending on the intercalators. The ethidium cation unwinds
Intercalation of planar organic compounds to DNA was first pro- DNA by about 26° and proflavine by about 17°. These structural
posed by Lerman to explain the strong affinity of certain heterocy- modifications can lead to functional changes, often leading to inhi-
clic aromatic dyes such as acridines for DNA [16]. Planar bition of transcription and replication and DNA repair processes,
heterocyclic compounds act as intercalators which stack between which make intercalators potent mutagens [23,24].
adjacent DNA base pairs leading to significant p-electron overlap. Intercalation is usually independent of the DNA sequence con-
Intercalators are molecules that stack perpendicular to the DNA text (a slight GC specificity has been observed). This mode of bind-
backbone without forming covalent bonds and without breaking ing is usually favored by the presence of an extended fused
up the hydrogen bonds between the DNA bases. The only known aromatic ligand as PHEHAT [25] or DPPZ [26]. With less extended
forces that sustain the stability of the DNA–intercalator complex, aromatic systems, the intercalation is usually prevented through
even more than DNA alone, are van der Waals, hydrogen bonding, clashing of the ancillary ligands with the phosphodiester backbone,
hydrophobic, and/or charge transfer forces [17–21]. DNA intercala- so that only partial intercalation can occur as it is the case for
M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19 5

Table 1
Examples of DNA-interacting agents.

Non-covalent binding agents Covalent DNA-adducts and direct DNA damage


Groove binding DNA intercalators
agents
Berenil Aminoacridines Busulfan
Bisbenzimadoles Arylaminoalcohols Camptothecin
Bleomycin Coumarins Chlorambucil
Chloroquine Cystodytin J Cis-platinum
Chromomycin A3 Diplamine Clomesone
Diamidine-2- YO-1 and YOYO-1 Cyclodisone
phenylindole
Distamycin A Daunomycin Ionizing radiation
Guanyl Quinolines and Nitrogen mustard
bisfuramidine quinoxalines
Mithramycin Ethidium bromide Nitrosoureas: 1,3-bis (2-chloroethyl)-1-nitrosourea; 1-(2-chloroethyl)-3-cyclohexyl-nitrosourea; 1-(2-chloroethyl)3-
Netamycin Proflavine (2,6-dioxo-3-piperidyl)-1-nitrosourea
Netropsin Echinomycin
Pentamidine Chlorpheniramine
Pilcamycin Methapyrilene
SN6999 Tamoxifen
SN7167 Bis-naphthalimide Oxidative-stress agents
Hoechst 33258 Doxorubicin UV radiation
M-AMSA Busulfan
Indoles Camptothecin
Cis-platinum

Scheme 2. (A) Cisplatin covalently bonded to DNA. B. (a) Modes of binding of cisplatin to guanine (G) and adenine (A); (b) 1,2-intrastrand GpG (structure a), 1,2-intrastrand
ApG (structure b), 1,3-intrastrand GpNpG (structure c), and 1,2-interstrand GpG (structure d).

[Ru(phen)3]2+ [25,26] (Fig. 2A and B) whereas Fig. 2C shows the grooves contribute to the complex stability of the threading inter-
intercalation of Co and Cu complexes into the DNA base pairs calators [32]. A threading intercalator occupies and interacts
[27]. Intercalation of a planar ligand of the complex in the DNA strongly with both the minor and major grooves of DNA simulta-
base pairs stack is shown in Fig. 2A [28,29]. neously. For example, the threading intercalation of acridine-4-
carboxamides into the duplex 50 -d(CG(5-BrU)ACG)2-30 (Fig. 3B)
4.2.1.1. Modes of intercalation. There are two major modes of inter- [33]. Acridine compounds are able to inhibit topoisomerase I and
calation: classical intercalation and threading intercalation [9,30]. II enzymes (topoisomerases are essential DNA-targeting enzymes
4.2.1.1.1. Classical intercalation. Classical intercalators, such as ben- that initially induce a DNA strand cleavage), render DNA damage,
zo[a]pyrene (BP) (Fig. 3A) [31], bind to DNA duplexes with essen- disrupt DNA repair and replication, and induce cell death [34].
tially all of their aromatic system inserted between GpG base
pairs that form the top and bottom of the intercalation site. 4.2.2. Groove binding
4.2.1.1.2. Threading intercalation. Threading intercalators usually Some small compounds bind to the minor groove of DNA by van
have two side chains on opposite sides of a planar aromatic ring der Walls interaction and hydrogen bonding. Minor groove binding
system, the process of complex formation with DNA is more com- drugs typically have several aromatic rings, such as pyrrole, furan
plicated. In such cases, one of the side-chains must slide through or benzene connected by bonds possessing torsional freedom. Min-
the intercalation cavity in order to form the complex. Favorable or groove binding drugs are usually narrow curved shaped, which
interactions of the side-chains with both the major and minor is isohelical to the curve of the minor groove, and facilitates
6 M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19

Scheme 3. Proposed mechanisms for the alkylation of N3 guanine by (+)-CC1065: (a) tautomerization of the G–C base pair to give the enol-G–imino-C tautomer; (b) a
concerted mechanism.

Scheme 4. Structures of the seven small molecules studied (P1–P7).

binding by promoting van der Waals interactions. Additionally, negative pockets of AT sequences is probably due to better van
these drugs can form hydrogen bonds to bases, typically to N3 of der Waals contacts between the ligand and groove walls in this re-
adenine and O2 of thymine. The groove-binding molecules are gion, since AT regions are narrower than GC groove regions and
commonly specific to adenine–thymine (AT) rich sequences. This also because of the steric hindrance in the latter, presented by
preference in addition to the designed propensity for the electro- the C2 amino group of the guanine base [35]. However, a few
M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19 7

Fig. 2. (A) Intercalation of a planar ligand of the complex in the DNA base pairs stack. (B) Intercalated ruthenium complex, [Ru(L)2(L0 )]2+ (L = bipyridine, 1,10-phenanthroline
(phen) and 1,4,5,8-tetraazaphenanthrene (tap) and L0 = dipyridophenazine (dppz) in double-stranded DNA fragment (CG–GC). (C) Intercalation of Co(II) and Cu(II) metal
complexes of N1,N5-bis[pyridine-2-methylene]-thiocarbohydrazone.

synthetic polyamides like lexitropsins and imidazole-pyrrole


polyamides have been designed which have specificity for GC
and CG regions in the grooves. Hydrophobic and/or hydrogen-
bonding are usually important components of this binding process,
and provide stabilization. The antibiotic netropsin is a model
groove-binder (Fig. 4A). Fig. 4B shows the generally proposed bind-
ing of acridine bis-imidazolidinones (R = ethyl) into adjacent Cyto-
sine: Guanine base pairs from minor groove side [36]. Geometric
and steric factors also play a role as shown with [Ru(TMP)3]2+,
where the methyl groups prevent intercalation [37] (Fig. 4C). Un-
like to intercalators, groove-binding drugs induce little or no struc-
tural rearrangement of the DNA helix. Fig. 4D shows the Groove
binding of Ni(II) and Zn(II) metal complexes of N1,N5-bis[pyri-
dine-2-methylene]-thiocarbohydrazone [28].

4.2.3. External binding


This type of binding is electrostatic in nature. Some ligands are
also capable of forming non-specific, outside edge stacking interac-
tions with the DNA phosphate backbone. This mode usually occurs
where the ligand self-associates to form higher-order aggregates,
which may stack on the anionic DNA backbone in order to reduce
charge–charge repulsion between ligand molecules. Some metal
complexes also interact with DNA through external binding, e.g.,
binding of Ru(II) complexes which are 2+ positive charge with Fig. 3. (A) NMR solution structure of the d(C4A5C6)-d(G13G14G15) segment of the
the DNA phosphate sugar backbone which is negatively charged. 10S-[BP]dAdG 9-mer duplex (50 -G1G2T3C4-[BP]A5C6G7A8G9-30 )/(50 -C10T11C12G13G14-
This association mode was proposed for [Ru(bpy)3]2+ as the lumi- G15A16C17C18-30 ), where the BP group attached to the A5 base intercalates
nescence enhancement of this complex upon binding to DNA is classically between the G13G14base step of the opposite strand. (B). X-ray structure
showing threading intercalation of 9-amino-6-bromo-DACA (space-filling model)
strongly dependent on the ionic strength. Cations like Mg2+, usu- into the DNA duplex 50 -d(C|G(5-BrU)AC|G)2-30 (wireframe), where | indicates
ally also interacts in this way (Fig. 5) [38]. intercalation sites.

5. Techniques used to study Drug–DNA interactions brief theoretical introduction to how it is applied to the study of
the interactions, followed by a discussion of some examples that
This section consists of a list of the different instrumental tech- are characteristic of the technique or that are of specific interest.
niques that have been used to study interactions between DNA and Various techniques that are used to study the binding of drug
small ligand molecules. For each technique, we have provided a molecules with DNA includes Infrared (IR), Raman, Circular
8 M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19

Fig. 4. (A) DNA complexed with netropsin, a minor groove binder. (B) Putative binding of acridine bis-imidazolidinones (R = ethyl) into adjacent cytosine:guanine base pairs
from minor groove side. (C) Adsorption of the [Ru(TMP)3]2+ in the DNA grooves. (D) Groove binding of Ni(II) and Zn(II) metal complexes of N1,N5-bis[pyridine-2-methylene]-
thiocarbohydrazone.

Fig. 5. External association of the complex in the atmosphere of ions of the DNA polyelectrolyte.

dichroism, UV–Visible, Nuclear magnetic resonance (NMR) spec- sumed that the magnitude of this shifting could be interpreted as
troscopies, Atomic force microscopy (AFM), Electrophoresis, Mass an indication of the strength of the interaction between the DNA
spectrometry, Viscosity measurements, Thermal denaturation structure and the ligand considered [39–42].
studies, Cyclic, Square wave and Differential pulse voltammetry, The UV–Visible absorption spectrum of DNA shows a broad
etc. These techniques have been used as a major tool to character- band (200–350 nm) in the UV region with a maximum absorption
ize the nature of drug–DNA complexation and the effects of such at 260 nm. This maximum is a consequence of the chromophoric
interaction on the structure of DNA. In this review article we will groups in purine (adenine and guanine) and pyrimidine (cytosine
focus on UV–Visible, fluorescence spectroscopies and cyclic vol- and thymine) moieties responsible for the electronic transitions.
tammetry as these are of common use. The probability of these transitions is high and thus the molar
absorptivity (e) is of order of 104 M1 cm1 [3]. This technique al-
5.1. UV–Visible absorption spectroscopy lows determining the molar concentration of DNA on the basis of
the measurement of the absorbance value at 260 nm. The absor-
UV–Visible absorption spectroscopy is perhaps the simplest and bance ratios (A260/A280 and A260/A230) can also describe the purity
most commonly employed instrumental technique for studying of DNA. This ratio should be in the range of 1.8–1.9 to ensure that
both the stability of DNA and their interactions with small ligand DNA is sufficiently free of protein [43]. Slight changes in the
molecules. The study of drug–DNA interactions could be carried absorption maximum as well as the molar absorptivity can occur
out by UV–Visible absorption spectroscopy by monitoring the with the variations in pH or ionic strength of the media. Drug–
changes in the absorption properties of the drug or the DNA mol- DNA interactions can be studied by comparison of UV–Visible
ecules. Usually, molecules used as ligands show an absorption absorption spectra of the free drug and drug–DNA complexes,
band that can be clearly distinguished in the visible region. An easy which are usually different. Compounds binding with DNA through
way to determine whether there is any interaction between the intercalation usually results in hypochromism and bathochromism
DNA and the drug is therefore to examine the shifting of the posi- (red shift). Because of the intercalative mode involving a stacking
tion of the maximum of this band from when the ligand is free in interaction between an aromatic chromophore and the base pair
solution to when the ligand is bound with the DNA. It has been as- of DNA, the extent of the hypochromism is usually consistent with
M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19 9

the strength of intercalative interaction [44]. The strength of this bance of DNA solution because many bases are in free form and
electronic interaction is expected to decrease as the cube of the do not form hydrogen bonds with complementary bases. As a re-
distance between the chromophore and the DNA bases decreases. sult, the absorbance for single-stranded DNA will be 40% higher
By decreasing the distance between intercalated compound (drug) than that for double stranded DNA at the same concentration. Fur-
and DNA bases, hypochromism take place apparently. Thus, this is thermore, the hyperchromic effect arises mainly due to the pres-
consistent with the combination of compound p electrons and p ence of charged cations which bind to DNA via electrostatic
electrons of DNA bases. Consequently, the energy level of the p– attraction to the phosphate group of DNA backbone and thereby
p electron transition decreases, which causes a red shift. This con- causing a contraction and overall damage to the secondary struc-
tributes to the hypochromic effect discussed above [45,46]. In case ture of DNA [47]. The hyperchromic effect may also be attributed
of electrostatic attraction between the compound and DNA, hyper- to external contact (electrostatic binding [48]) or to partial uncoil-
chromic effect is observed that reflects the corresponding changes ing of the helix structure of DNA, exposing more bases of the DNA
of DNA in its conformation and structure after the complex–DNA [49]. If there is a weaker interaction then only hypochromic or
interaction has occurred. The hyperchromic effect is the outstand- hyperchromic effects are observed without significant changes of
ing increase in absorbance of DNA upon denaturation. The two shifts in the spectral profiles [3,50].
strands of DNA are held together mainly by the stacking interac- Based upon the variation in absorbance, the intrinsic binding
tions, hydrogen bonds and hydrophobic effect between the com- constant/association constant (K) of the drug with DNA can be
plementary bases. The hydrogen bond limits the resonance of the determined according to Benesi–Hildebrand equation [43]:
aromatic ring so the absorbance of the sample is limited as well.
A0 eG eG 1
When the DNA double helix is treated with denaturing agents, ¼ þ 
the interaction force holding the double helical structure is dis-
A  A0 eHG  eG eHG  eG K½DNA
rupted. The double helix then separates into two single strands where K is the association/binding constant, A0 and A are the absor-
which are in the random coiled conformation. At this time, the bances of the drug and its complex with DNA, respectively, and eG
base–base interaction will be reduced, increasing the UV absor- and eH–G are the absorption coefficients of the drug and the

Fig. 6. (A) Absorption spectra of [{SnPh3(O2CCH2SXyl)}1, where Xyl = 3,5-Me2C6H3] in the presence of increasing amounts of DNA. Arrow indicates that absorbance changes
upon increasing DNA concentrations. Inset: plot of [DNA]/eo  ef = [DNA]/eo  ef + 1/Kb(/eo  ef). (B) Electronic spectral of [Cu(L)(H2O)2](NO3)2 (L = 2,6-bis(benzimidazol-
yl)pyridine) through titration with CT-DNA in tris–HCl buffer; C. Plot of [DNA]/(eo  ef) vs. [DNA] for the titration of CT-DNA with [Cu(L)(H2O)2](NO3)2 for the determination of
Kb (0.8  105 M1).
10 M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19

Fig. 7. UV–Visible absorption spectra of complexes C10H22N2O5SnCl2 (A) and C10H22N2O5ZrCl2 (B) in Tris–HCl buffer upon addition of CT-DNA. Plots of [DNA] vs. [DNA]/ea  ef
for the titration of CT-DNA with complexes C10H22N2O5SnCl2 (C) and C10H22N2O5ZrCl2 (D).

drug–DNA complex, respectively. The association constant can be ob-


tained from the intercept-to-slope ratios of A0/(A  A0) vs. 1/[DNA]
plots. The K can also be determined from the intercept-to-slope ratios
of the plot of [DNA] vs. [DNA]/ea  ef, where ea (or eG) and ef (or eH–G)
are the absorption coefficients of the drug and the drug–DNA com-
plex, respectively.
Fig. 6A, describes the interaction of triphenyltin(IV) carboxylate
complex [{SnPh3(O2CCH2SXyl)}1, where (Xyl = 3,5-Me2C6H3)] with
DNA. The inset graph represents the plot of [DNA]/eo  ef = [DNA]/
eo  ef + 1/Kb(/eo  ef), used for the determination of the intrinsic
binding constant, Kb (Kb = 1.68  105 M1) [51]. As it is shown in
Fig. 6A, with increasing concentrations of DNA, the absorption
bands of the complex were affected, resulting in the tendency of
hyperchromism and a very slight blue shift. The complex [{SnPh3(-
O2CCH2SXyl)}1] may be charged (R3Sn+), The R3Sn+ moieties are
assumed to directly affect DNA [52] as well as binding to mem-
brane proteins or glycoproteins, or to cellular proteins, and there
could be classical electrostatic interactions with the negatively
charged oxygen of the phosphate group of DNA, which may be
responsible for the spectral changes observed. However, other
electrostatic effects such as hydrogen bonding between the com-
plexes and the base pairs in DNA may also be present [44,53,54].
Fig. 6B represent the interaction of copper(II) complex with Fig. 8. Absorption spectra of Zn(Pyimpy)2](ClO4)2C6H5CH30.5H2O in 0.1 M phos-
DNA. With increase in DNA concentration (indicated by an arrow) phate buffer (pH 7.2) containing 5% DMF in the presence of increasing amounts of
DNA. Arrows show the absorbance changes upon increasing DNA concentration.
M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19 11

Fig. 9. Absorption spectra of 0.2 mM Ph2SnL (A), Me2SnL (B) and Bu2SnL (C) in the absence and presence of 5–25 mM DNA. The arrow direction indicates increasing
concentrations of DNA. (D). Plots of A0/(A – A0) vs. 1/[DNA] for the determination of binding constants of Complex-DNA adducts. Absorption spectra of 3 mM Bu3SnL (E),
Cy3SnL (F) and Ph3SnL (G) in the absence and presence of 20 lM, (b) 30 lM, (c) 40 lM, (d) 50 lM, (e), 60 lM (f), 70 lM (g) and 80 lM DNA (h) in 10% aqueous DMSO at 25 °C.
The arrow direction indicates increasing concentrations of DNA.

hyperchromic effect is observed. A significant hyperchromism ef- structure after the complex–DNA interaction has occurred. Fur-
fect centered at the 330 nm absorption maximum, suggest a strong thermore, the hyperchromic effect arises mainly due to the pres-
interaction between the copper(II) complex and DNA. This spectral ence of charged cations Sn(IV)/Zr(IV) which bind to DNA via
change might be indicative of groove binding. Groove binding mol- electrostatic attraction to the phosphate group of DNA backbone
ecules contain unfused aromatic ring systems connected by bonds and thereby causing a contraction and overall damage to the sec-
with torsional freedom in order to adopt appropriate conformation ondary structure of DNA [47].
that closely matches the helical turn of DNA grooves, while fused Fig. 8 describes the interaction of Zn complex with CT-DNA. A
polyaromatic systems containing protonated nitrogen atoms or blue shift of 4 nm and hyperchromic shift of the absorption band
having protonated side chains attached to the ring system are typ- were observed in case of complex Zn(Pyimpy)2](ClO4)2C6H5CH3-
ically intercalators. Here, the organic ligand L, having coplanar 0.5H2O. Isosbestic point near 350 nm was observed for the com-
atoms upon copper coordination, likely facilitates the formation plex. Such type of spectral changes describes the covalent
of Van der Waals contacts or hydrogen bonds during interaction binding of zinc complex with DNA. Zinc complexes in the presence
with DNA grooves [55]. of DNA get attached to the nucleic acid base(s) and a new species
The complexes C10H22N2O5SnCl2 (1) and C10H22N2O5ZrCl2 (2) generates during DNA interaction. The intrinsic binding constant K
exhibit intraligand absorption bands in the UV region at 248 and for the complex was found to be 2.86  103 M–1. The observed
253 nm, respectively, as shown in Fig. 7A and B. By the addition binding constant is smaller than the classical intercalators and
of increasing amounts of CT-DNA ((0.066–0.333)  104 M) to metallointercalators where binding constant was reported to be
complexes C10H22N2O5SnCl2 and C10H22N2O5ZrCl2, a sharp hyper- in the order of 107 M–1[56,57].
chromic effect in the absorption bands with a moderate red shift Fig. 9 describes the effect of varying concentration of DNA (5–
of 5 and 3 nm, respectively is observed. Hyperchromic effect re- 25 mM) on the electronic absorption spectra of 0.2 mM of Me2SnL
flects the corresponding changes of DNA in its conformation and (A), Bu2SnL (B) and Ph2SnL (C) (L = N0 -(2-hydroxybenzylidene)
12 M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19

A similar intercalation mode of interaction of Bu3SnL (E), Cy3SnL


(F) and Ph3SnL (G) (L = 4-(4-nitrophenyl)piperazinium 4-(4-nitro-
phenyl)piperazine-1-carbodithioate) with DNA is shown in
Fig. 9E–G. The binding of complexes Bu3SnL, Cy3SnL and Ph3SnL
to DNA caused a progressive blue shift of 10 (401–391), 8 (400–
392) and 4 nm (398–394), respectively. Such spectral characteris-
tics are indicative of their binding to DNA. The peculiar hypochro-
mism observed here is attributed to the intercalation of these
drugs into the DNA base pairs. The binding constant value deter-
mined for these compounds are 6.05  103, 2.30  103 and
3.25  103 M1, respectively. The greater value of K for the Bu3SnL
is due to additional hydrophobic nature of butyl group with the
bases of DNA [60].
Absorption spectrum of GMB exhibited bands at 207, 232, 248,
283 and 331 nm (Fig. 10). With the addition of increasing amounts
of DNA the absorbance of GMB increased at 207, 232, 248 and 283
(with red shift at 207 and 248 nm), while those at 331 nm de-
creased slightly. These results revealed that there is strong interac-
Fig. 10. UV–Visible absorption spectra of 2.5  105 M GMB in presence of different
tion between drug and DNA. The appreciable shift in wavelength
concentrations of DNA 0 (1), 7.5 (2), 15 (3), 22.5 (4), 30 (5), 37.5 (6), 45 (7), 52.5 and
(8) 60 lM. (red shift P 15 nm) indicates the intercalative binding mode of
molecules to DNA helix due to the interaction of a DNA p stack
formohydrazide. The strong absorption of these compounds in the with ligand, while outside binders (groove binders) shows a smal-
near UV region (292–330 nm) is due to the long-living triplet ex- ler red shift (Dk 6 8 nm). The red shift for absorption peaks at 207
cited state of the aromatic system. The broad absorption bands in and 248 is about 1 and 5 nm, respectively, so groove binding mode
the region (350–470 nm) is due the intraligand transition between between GMB and DNA is also there. The binding constant, K, cal-
p and p energy levels of the tridentate ligand N0 -(2-hydroxyben- culated is 1.97  106 M1[61].
zylidene)formohydrazide. The absorption spectra of Me2SnL, Bu2- Fig. 11 shows the effect of different concentration of DNA (10–
SnL and Ph2SnL recorded a 68.26%, 62.83% and 65.91% decrease 60 lM) on the electronic absorption spectrum of 30 lM NFC. The
in peak intensities accompanied with 2, 5 and 3 nm red shift rationale behind the band in the UV region (328 nm) is the proba-
respectively, by the addition of 25 mM DNA. The peculiar hypo- ble charge transfer between the non-bonding or antibonding orbi-
chromic effects can be associated with the interaction of electronic tal of the cyclopentadienyl ring and the iron atom of NFC. The
states of the intercalating chromophore and those of the stacked maximum absorption of the drug at this wavelength exhibited
base pairs of DNA. The slight bathochromic shifts can best be de- slight bathochromic and pronounce hypochromic shifts by the
scribed by the lowering in p and p transition energy of the ligands incremental addition of DNA. The bathochromic effect is associated
in diorganotin(IV) complexes due to their ordered stacking be- with the decrease in the energy gap between the highest (HUMO)
tween the DNA base pairs after intercalation. After binding to the and the lowest molecular orbitals (LUMO) after the interaction of
DNA, the p orbital of the binding ligand could couple with p orbi- NFC to DNA. The compactness in the structure of either the drug
tal of base pairs in the DNA. The coupling p orbital was partially alone and/or DNA after the formation of drug–DNA complex may
filled by electrons, thus decreasing the transition probabilities, result in hypochromism [62].
and hence resulting in the hypochromicity. The binding constants,
with values of 1.54  104, 8.19  103 and 2.59  104 M1 for the 5.2. Fluorescence emission spectroscopy
interaction of Me2SnL, Bu2SnL and Ph2SnL with DNA were obtained
from the slope to intercept ratio of the plots (D) between Fluorescence spectroscopy is probably one of the techniques
A0/(A  A0) vs. 1/[DNA] [58,59]. most commonly used to study interactions between small ligand
molecules and DNA. The advantages of molecular fluorescence
over other techniques are its high sensitivity, large linear concen-
tration range and selectivity. The most intense and the most useful
fluorescence is found in compounds containing aromatic func-
tional groups with low-energy p ? p transition levels. Com-
pounds containing aliphatic and alicyclic carbonyl structures or
highly conjugated double-bond structures may also show fluores-
cence, but the number of these transitions is small compared with
those in aromatic systems [40,63].
The mode of binding of drugs to DNA can be determined by
fluorescence spectroscopy and the various analytical tools based
on fluorescence emission can also provide particularly useful infor-
mation. The orientation of fluorophoric ligands and their closeness
to the DNA pairs of bases can be studied by fluorescence anisotropy
or fluorescence resonance energy transfer.
Fluorescence quenching experiments give additional informa-
tion concerning the localization of the drugs and their mode of
interaction with DNA [3].
Fluorescence emission is very sensitive to the environment, and
hence the fluorophore transfer from high to low polarity environ-
Fig. 11. UV–Visible absorption spectra of 30 lM NFC in the absence of DNA (a) and ments usually causes spectral shifts (10–20 nm) in the excitation
presence of 10–60 lM DNA (b–g) in 10% aqueous ethanol buffered at pH 6. and emission spectra of drugs [64]. Moreover, the effective
M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19 13

Fig. 12. Emission spectra of EB bound to CT-DNA in the presence of the Mn complex [C42 H40 Mn N14 O16] (a) and Zn complex [C78 H62 N26 O28 Zn2] (c). The arrows show the
intensity changes upon increasing concentrations of the complexes. Fluorescence quenching curves of EB bound to CT-DNA by the Mn complex (b) and Zn complex (d). Plots
of I0/I vs. [Complex].

interaction with DNA usually causes a significant enhancement served in the presence of quencher. KSV can be obtained from the
of the fluorescence intensity as a consequence of different fac- slope of the plot of Fo/Fvs. [DNA] [3,66].
tors. Thus, in the case of intercalating drugs, the molecules are Ethidium bromide (EB) is a common fluorophore that bind to
inserted into the base stack of the helix. The rotation of the free DNA. The fluorescene of EB increases in the presence of CT-DNA,
molecules favors the radiationless deactivation of the excited due to its strong intercalation between the adjacent CT-DNA base
states, but if the drugs are bound to DNA the deactivation pairs. It was previously reported that the enhanced fluorescence
through fluorescence emission is favored, and a significant in- can be quenched by the addition of a second molecule. Thus if
crease in the fluorescence emission is normally observed. In case the second molecule intercalates into DNA, it leads to a decrease
of groove binding agents, electrostatic, hydrogen bonding or in the fluorescence intensity of the EB-DNA because it will compete
hydrophobic interactions are involved and the molecules are with EB in binding with DNA. The extent of fluorescence quenching
close to the sugar-phosphate backbone, being possible to observe of EB bound to CT-DNA can be used to determine the extent of
a decrease in the fluorescence intensity in the presence of DNA binding between the second molecule and CT-DNA [66–68]. The
[65]. The use of well-established quenchers, i.e., halide ions, pro- emission spectra of EB bound to DNA in the absence and presence
vides further information about the binding of drugs to DNA. The of complex are shown in Fig. 12a–d, where the fluorescence inten-
groove binders are more sensitive to the quenching effect by ha- sity of EB-DNA decreases in the presence of a second molecule that
lides than the intercalating agents, because the pairs of bases intercalates itself into the base pair of the DNA. The fluorescence
hinder the accessibility of the drug by the quenchers. Besides, quenching of EB bound to CT-DNA by Mn complexe [C42 H40 Mn
the electrostatic repulsive forces among phosphate groups on N14 O16] and Zn complex [C78 H62 N26 O28 Zn2] is shown in
DNA and anionic quenchers collaborate to protect the drug from Fig. 12a and c. The quenching of EB bound to CT-DNA by both com-
the quencher effects. Thus, in the case of intercalating agents plexes is in good agreement with the linear Stern–Volmer equa-
reduction in the KSV {Stern–Volmer equation, Fo/F (or I0/ tion, which provides further evidence that the complexes bind to
I) = 1 + KSV [Q], where Fo or I0 is the fluorescence intensity in DNA. The KSV values for Mn and Zn complexes and are
the absence of quencher while F or I is the fluorescence intensity 2.9  103 M1 and 1.9  103 M1, respectively. The data suggest
in the presence of quencher, KSV is the Stern–Volmer quenching that the interaction of Mn complex with CT-DNA is stronger than
constant, [Q] is the concentration of the quencher} values is ob- that of Zn complex [67].
14 M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19

Fig. 13. Emission spectra of EB bound to CT-DNA in the presence of free HL [N-(2-hydroxylacetophenone)-3-oxapentane-1,5-diamine] (A) and Ni complex [Ni2(L)2(NO3)2] (C),
kex = 520 nm. The arrows show the intensity changes upon increasing concentrations of the complexes. Fluorescence quenching curves of EB bound to CT-DNA by the free HL
(B) and Ni complex (D). (Plots of I0/I vs. [Compound]).

Similarly Fig. 13 describes the interaction of N-(2-hydroxylace- interactions by the reliance of the current passed during oxidation
tophenone)-3-oxapentane-1,5-diamine (HL) and its Ni complex or reduction of the bound species on the amount of added DNA. In
with EB-DNA. The KSV values of compound are (1.7 ± 0.105) some cases it should also be possible to obtain kinetic data from
9  103 M1 and (7.0 ± 0.097)  103 M1 for HL and the complex, current and potential measurements, since voltammetric methods
respectively. These values show that both the free ligand and com- are sensitive to chemical reactions (e.g., ligand dissociation) cou-
plex can compete for DNA binding sites and so displace EB from the pled to the electron-transfer steps [70].
DNA, which is usually characteristic of the intercalative interaction Cyclic voltammetry (CV) is widely used for the evaluation of mode
of compounds with DNA. Fig. 14 describe the interaction of DNA of action and binding strength of drug–DNA interaction. This technique
with copper complexes [55,68,69]. is predominantly useful for metal based compounds due to their acces-
sible redox states. As in CV the scan is revered so the fate of the species
5.3. Cyclic voltammetry in the backward scan can also be studied. The peak potential and peak
current of the compound changes in the presence of DNA if the com-
The application of electrochemical methods to the study of met- pound interacts with it. The variation in peak potential and peak cur-
allointercalation and coordination of metal ions and chelates to rent can be exploited for the determination of binding parameters.
DNA provides a useful complement to the previously used meth- The decay in peak current (Ip) of the drug by the addition of increasing
ods of investigation, such as UV–Visible spectroscopy. Small mole- amount of DNA can be used for the determination of binding constant
cules which are not amenable to such methods, either because of and binding site size, whereas the shift in peak potential can be used to
weak absorption bands or because of overlap of electronic transi- ascertain the mode of interaction. The binding constant is quantified
tions with those of the DNA molecule, can potentially be studied by the following equation [71]:
via voltammetric techniques. Multiple oxidation states of the same
logð1=½DNAÞ ¼ log K þ logðI=ðI0  IÞÞ
species as well as mixtures of several interacting species can be ob-
served simultaneously. Equilibrium constants (K) for the interac- where K is the binding constant, I0 and I are the peak currents of the
tion of the metal complexes with DNA can be obtained from drug in the absence and presence of DNA, respectively. The binding
shifts in peak potentials, and the number of base pair sites involved constant, K is obtained from the intercept of the plot of log (1/
in binding(s) through intercalative, electrostatic, or hydrophobic [DNA]) vs. log (I/(I0  I)).
M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19 15

Fig. 14. (A) Emission spectrum of EtBr bound to DNA in the presence of [Cu2L(phen)](ClO4)2 where L = 6,60 -piprazine-1,4-diyldimethylene bis (4-methyl phenol). The arrow
shows the intensity change upon increasing Cu(II) complex (3) concentrations. Inset shows the plots of emission intensity I0/I vs. [DNA]/[complex for determining KSV. (B).
Emission spectra of the CT-DNA–EB system in tris–HCl buffer in the presence of [Cu(L)(H2O)2](NO3)2, where L = 2,6-bis(benzimidazolyl)pyridine; kex = 522 nm. The arrow
indicates the increase of the complex concentration. (C). Plot of I0/I vs. [complex] for the titration of CT-DNA–EB system with [Cu(L)(H2O)2](NO3)2 for determining KSV
(10.0  104). (D) Plot of Io/I vs. [complex] for the titration of CT-DNA–EB system with 1 using spectrofluorimeter; linear Stern–Volmer quenching constant (Ksv) for 1 =
10.0104; (R = 0.99596 for five points).

Another equation which can be used for the determination of


binding constant is given as [72]:
1 Kð1  AÞ
¼ K
½DNA 1  I=I0
where A is proportionality constant. The binding constant is calcu-
lated from the intercept of the plot of 1/[DNA] vs. 1/(1  I/I0).
A simple site binding model used to fit the cyclic voltammetric
data acquired from the interaction between the drug and DNA is
given as [73]:
C b =C f ¼ Kf½free base pairs=sg
where s is the binding site size in terms of base pairs, Cf is the con-
centration of the free species, Cb denotes the concentration of DNA-
bound species and K represents binding constant.
Measuring the concentration of DNA in terms of nucleotide
phosphate, the concentration of the base pairs can be expressed
as [DNA]/2 and the above equation can be written as
C b =C f ¼ Kf½DNA=2sg
The Cb/Cf was determined by the equation given below [74]:
Fig. 15. Square wave voltammograms for 0.09 mM quinacrine in the ‘‘a’’ absence
C b =C f ¼ ðI0  IÞ=I and ‘‘b’’ presence of 0.5 mM DNA in 100 mM Tris–HCl buffer at pH 7.0. Equilibrium
time: 10 s, frequency: 10 Hz, step potential: 5 mV, amplitude: 25 mV.
The right hand side of the above equation can be obtained di-
rectly from the experimental peak currents. The experimental data
can therefore be expressed as plots of Cb/Cf against the total con- It has been reported in the literature that the binding constants
centration of DNA keeping concentration of the drug constant. for redox active species can be obtained from straightforward
16 M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19

Carter and Bard have reported three kinds of modes for small
molecules binding to DNA, if E1/2 shifted to more negative value
when small molecules interacted with DNA; the interaction mode
was electrostatic binding. On the other hand, if E1/2 shifted to more
positive value, the interaction mode was intercalative binding [70].
Fig. 15 shows the effect of the addition of DNA to a solution of
0.09 mM quinacrine in 100 mM Tris–HCl buffer, pH 7, on the
square wave voltammetry of quinacrine. The current drops on
the addition of DNA owing to the binding of quinacrine. Further-
more, the peak potential shifted to a more positive value in the
presence of DNA. The shift in peak potential is typical of the inter-
calation of species into DNA [21,30]. In the presence of nucleic
acids, the current is mainly due to free species, as the diffusion rate
of bound species is small [26]. The cause for the decrease in the
peak current was that the obvious diffusion coefficient and the
obvious concentration of electroactive species decreased [75,77].
Fig. 16 describes the cyclic voltammetric behavior of dibutyl-
tin(IV) bis(4-nitrophenylethanoate) on glassy carbon electrode in
the absence and presence of DNA, in 10% aqueous DMSO at
25 °C. Due to the instability of Sn3+, the voltammetric response is
attributed to the 2e redox process of Sn2+/Sn4+ couple. In the ab-
sence of DNA, dibutyltin(IV) bis(4-nitrophenylethanoate) regis-
Fig. 16. CV behavior of 3 mM dibutyltin(IV) bis(4-nitrophenylethanoate) at GC
tered a cathodic peak at 1.212 V and anodic peak at 1.014 V.
electrode in the absence (a) and presence of 10 (b), 20 (c), 30 (d), 40 (e), 50 (f) and
60 lMDNA (g) in 10% aqueous DMSO with 0.1 M TBAFB as supporting electrolyte at The large peak to peak potential difference (DEp) of 198 mV is sug-
0.1 V/s scan rate. gestive of electrochemical reaction coupled with a chemical reac-
tion. With the increase in concentration of DNA in a constant
amount of dibutyltin(IV) bis(4-nitrophenylethanoate), the voltam-
voltammetric experiments in which the DNA is titrated against the metric response of the compound changed as is evidenced by the
redox active molecule. The measurement of diffusion currents in sequential drop in peak current and gradual peak potential shift
the presence of excess nucleic acid has shown that the diffusion in positive direction. The shift of peak potential to less negative
coefficient of DNA-bound species is more than one order of magni- values is suggestive of intercalation of dibutyltin(IV) bis(4-nitroph-
tude lower than that of the free species [74–76]. enylethanoate) into DNA. The observed decrease in peak current

Fig. 17. Cyclic voltammograms of 3 mM: (A) Ph2SnL (1), (B) Me2SnL (2) and (C) Bu2SnL (3) in 10% aqueous DMSO with 0.1 M TBAP as supporting electrolyte in the absence (a)
and presence of 50 mM DNA (b) at 25 °C. (D) Plots of log (IH–G/(IG  IH–G)) vs. log (1/[DNA]) used to calculate the binding constants of 1-DNA, 2-DNA and 3-DNA adducts. (E)
Plots of I vs. m1/2, for the determination of diffusion coefficients of the free drugs (3 mM 1–3). Scan rates 0.1–0.6 V/s with a difference of 0.1 V/s. (F) Plots of I vs. m1/2, for the
determination of the diffusion coefficients of the DNA bound drugs by taking 3 mM 1–3 and 60 mM DNA. Scan rate 0.1–0.6 V/s with a difference of 0.1 V/s.
M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19 17

Fig. 18. (A) Cyclic voltammograms of the copper(II) complex in the absence (a) and presence (b) of HS-DNA. (B) The plot of 1/[DNA] vs. 1/(1  i/i0).

indicates the formation of large and slowly diffusing dibutyltin(IV) In UV–Visible spectroscopy, hypochromism and red shift are the
bis(4-nitrophenylethanoate)-DNA adduct due to which the free signs of intercalation while hyperchromism is the sign of electro-
drug concentration (which is mainly responsible for the conduc- static mode of interaction. Similarly in cyclic voltammetry shift
tion of the current) is lowered [78]. of E1/2 towards more positive value is the indication of intercala-
Similarly Fig. 17 describes the intercalative mode of interaction tion mode of interaction while that towards more negative value
of organotin(IV) complexes with DNA. The shift in peak potential is the indication of electrostatic mode of interaction. In the case
toward positive side could be attributed to the intercalation of of fluorescence, a significant increase in the fluorescence emission
compounds A–C into the stacked base pair pockets of DNA [58]. is normally observed for intercalative mode of interaction while a
Fig. 18 describes the interaction of [Cu2(dmaeoxd)(ox)]nnH2O decrease in the fluorescence intensity is observed for groove bind-
with HS-DNA via groove binding. It can be seen from Fig. 18, in ing agents, electrostatic, hydrogen binding or hydrophobic
the absence of DNA (curve a), the copper(II) polymer has been interactions.
found to show a quasi-reversible redox process corresponding to
Cu(II)/Cu(I) with the cathodic (Epc) and anodic peak potential 7. Abbreviations
(Epa) being 0.181 and 0.097 V (DEp = 0.084 V), respectively.
The formal potential (Eo0 ) was found to be 0.139 V. In the pres-
ence of DNA (curve b) with R = 10 (R = [DNA]/[complex]) the vol- A adenine
tammetric peak currents decreased, suggesting that there exists T thymine
an interaction between the copper(II) polymer and HS-DNA. The G guanine
peak-to-peak separation becomes larger with DEp = 0.087 V, indi- C cytosine
cated that the electron-transfer process seems to become less FEBS Federation of European Biochemical Societies
reversible in the presence of the DNA. The ratio of the constants TMP 3,4,7,8-tetramethyl phenanthroline
for binding the Cu(I) and Cu(II) ions to HS-DNA is estimated to byp bipyridine
be 0.9–1, indicating that both Cu(I) and Cu(II) interact with HS- DPPZ dipyridophenazine
DNA almost to the same extent. Since the complex has no charged PHEHAT 1,10-phenanthrolino[5,6-b]-1,4,5,8,9,
groups and aromatic rings, therefore, it may bind to HS-DNA in a 12-hexaazatriphenylene
groove-binding mode [72]. EB ethidium bromide
CT-DNA calf thymus DNA
6. Conclusions HS-DNA herring sperm DNA
Tris tris(hydroxymethyl)aminomethane
The review article has focused on drug–DNA interactions and MLCT metal-to-ligand charge transfer
their study by various techniques like UV–Visible, fluorescence GMB gemcitabine hydrochloride
spectroscopies and cyclic voltammetry. These techniques are used Pyimpy (2-((2-phenyl-2-(pyridin-2-
to evaluate the binding mode as well as binding strength of the l)hydazono)methyl)pyridine)
complex formed between Drug and DNA. Basically small ligand Hnap naproxen
molecules (drug) interact with DNA via two different ways: cova- Nadicl sodium diclofenac
lent and non-covalent modes. Covalent binders act as alkylating NFC 4-nitrophenylferrocene
agents as they alkylate the nucleotides of DNA. The non-covalent dmaeoxd N,N0 -bis[2-(dimethylamino)ethyl]oxamide
binders interact with via three different ways: intercalation, ox oxalate
groove binding and external binding (on the outside of the helix).
18 M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19

[28] D. Ambrosek, P.F. Loos, X. Assfeld, C. Daniel, A theoretical study of Ru(II)


Acknowledgment polypyridyl DNA intercalators: structure and electronic absorption
spectroscopy of [Ru(phen)2(dppz)]2+ and [Ru(tap)2(dppz)]2+ complexes
The author is acknowledged the HEC Islamabad, Pakistan for intercalated in guanine–cytosine base pairs, J. Inorg. Biochem. 104 (2010)
893–901.
financial support. [29] A.D. Tiwari, A.K. Mishra, S.B. Mishra, B.B. Mamba, B. Maji, S. Bhattacharya,
Synthesis and DNA binding studies of Ni(II), Co(II), Cu(II) and Zn(II) metal
References complexes of N1, N5-bis[pyridine-2-methylene]-thiocarbohydrazone Schiff-
base ligand, Spectrochim. Acta A 79 (5) (2011) 1050–1056.
[30] L.G. Bulnes, J. Gallego, Indirect effects modulating the interaction between
[1] E.D. Gadsby, Drug–DNA Interactions and Condensations Investigation with
DNA and a cytotoxic bisnaphthalimide reveal a two-step binding process, J.
Atomic Force Microscopy, PhD Dissertation, School of Chemistry and
Am. Chem. Soc. 131 (2009) 7781–7791.
Biochemistry, Georgia Institute of Technology, July 2004, pp. 23–25.
[31] H.J.C. Yeh, J.M. Sayer, X. Liu, A.S. Altieri, R.A. Byrd, M.K. Lakshman, H. Yagi, E.J.
[2] M.M. Harding, C.J. Harden, L.D. Field, A P-31 NMR-study of the interaction of
Schurter, D.G. Gorenstein, D.M. Jerina, NMR solution structure of a
the antitumor active metallocene CP2MoCl2 with calf thymus DNA, FEBS Lett.
nonanucleotide duplex with a dG mismatch opposite a 10S adduct derived
322 (3) (1993) 291–294.
from trans addition of a deoxyadenosine N6-amino group to (+)-
[3] V. Gonzalez-Ruiz, A.I. Olives, M.A. Martin, P. Ribelles, M.T. Ramos, J.C.
(7R,8S,9S,10R)-7,8-dihydroxy-9,10-epoxy-7,8,9,10-
Menendez, An overview of analytical techniques employed to evidence
tetrahydrobenzo[a]pyrene: an unusual syn glycosidic torsion angle at the
drug–DNA interactions. Applications to the design of genosensors, in: M.A.
modified dA, Biochemistry 34 (1995) 13570–13581.
Komorowska, S. Olsztynska-Janus (Eds.), Biomedical Engineering, Trends,
[32] A. Paul, S. Bhattacharya, Chemistry and biology of DNA-binding small
Research and Technologies. In Tech, 2011, pp. 65–90.
molecules, Curr. Sci. 102 (2) (2012) 212–231.
[4] O. Kennard, DNA–drug interactions, Pure Appl. Chem. 65 (6) (1993) 1213–
[33] A.K. Todd, A. Adams, J.H. Thorpe, W.A. Denny, L.P.G. Wakelin, C.J. Cardin, Major
1222.
groove binding and a ‘‘DNA-induced’’ fit in the intercalation of a derivative of
[5] R. Hajian, N. Shams, M. Mohagheghian, Study on the interaction between
the mixed topoisomerase I/II inhibitor DACA into DNA; X-ray structure
doxorubicin and deoxyribonucleic acid with the use of methylene blue as a
complexed to d(CG(5BrU)ACG)2 at 1.3 Å resolution, J. Med. Chem. 42 (1999)
probe, J. Braz. Chem. Soc. 20 (8) (2009) 1399–1405.
536–540.
[6] S. Rauf, J.J. Gooding, K. Akhtar, M.A. Ghauri, M. Rahman, M.A. Anwar, A.M.
[34] L. Janovec, M. Kozurkova, D. Sabolova, J. Ungvarsky, H. Paulikova, J. Plsikova, Z.
Khalid, Electrochemical approach of anticancer drugs–DNA interaction, J.
Vantova, J. Imrich, Cytotoxic 3,6-bis((imidazolidinone)imino)acridines:
Pharmaceut. Biomed. Anal. 37 (2005) 205–217.
synthesis, DNA binding and molecular modeling, Bioorg. Med. Chem. 19
[7] R.D. Snyder, L.B. Hendry, Toward a greater appreciation of noncovalent
(2011) 1790–1801.
chemical/DNA interactions: application of biological and computational
[35] H.Y. Mei, J.K. Barton, Chiral probe for A-form helices of DNA and RNA-
approaches, Environ. Mol. Mutagen. 45 (2005) 100–105.
tris(tetramethylphenanthroline)ruthenium(II), J. Am. Chem. Soc. 108 (1986)
[8] C. Silvestri, J.S. Brodbelt, Tandem mass spectrometry for characterization of
7414–7416.
covalent adducts of DNA with anticancer therapeutics, Mass Spectrom Rev. 2
[36] M.L. Kopka, C. Yoon, D.S. Goodsell, P. Pjura, R.E. Dickerson, Binding of an
(2012) 1–20.
antitumor drug to DNA: netropsin and C-G-C-G-A-A-T-T-BrC-G-C-G, J. Mol.
[9] H.K. Liu, P.J. Sadler, Metal complexes as DNA intercalators, Acc. Chem. Res. 44
Biol. 183 (1985) 553–563.
(2011) 349–359.
[37] I. Haq, J. Ladbury, Drug–DNA recognition: energetics and implications for
[10] N. Hadjiliadis, E. Sletten, Metal complex–DNA interactions, Blackwell
design, J. Mol. Recognit. 13 (2000) 188–197.
Publishing Ltd., 2009. pp. 138–139.
[38] J.M. Kelly, A.B. Tossi, D.J. McConnell, C. Oh Uigin, A study of the interactions of
[11] ‘‘Antineop’’, <http://faculty.swosu.edu/scott.long/phcl/antineop.htm>
some polypyridylruthenium(II) complexes with DNA using fluorescence
(retrieved 24.01.09).
spectroscopy, topoisomerisation and thermal denaturation, Nucl. Acids Res.
[12] S.R. Rajski, R.M. Williams, DNA cross-linking agents as antitumor drugs, Chem.
13 (1985) 6017–6034.
Rev. 98 (1998) 2723–2796.
[39] H. Sun, J. Xiang, Y. Liu, L. Li, Q. Li, G. Xu, Y. Tang, A stabilizing and denaturing
[13] N. Kondo, A. Takahashi, K. Ono, T. Ohnishi, DNA damage induced by alkylating
dual-effect for natural polyamines interacting with G-quadruplexes depending
agents and repair pathways, J. Nucleic Acids 2010 (2010) 1–7.
on concentration, Biochimie 93 (2011) 1351–1356.
[14] H.J. Park, L.H. Hurley, Covalent modification of N3 of guanine by (+)-CC-1065
[40] J. Jaumot, R. Gargallo, Experimental methods for studying the interactions
results in protonation of the cross-strand cytosine, J. Am. Chem. Soc. 119
between G-quadruplex structures and ligands, Curr. Pharmaceut. Des. 18 (14)
(1997) 629–630.
(2012) 1900–1916.
[15] Y. Ni, D. Lin, S. Kokot, Synchronous fluorescence, UV–visible
[41] C. Wei, J. Wang, M. Zhang, Spectroscopic study on the binding of porphyrins to
spectrophotometric, and voltammetric studies of the competitive interaction
(G4T4G4)4 parallel G-quadruplex, Biophys. Chem. 148 (2010) 51–55.
of bis(1,10-phenanthroline)copper(II) complex and neutral red with DNA,
[42] K. Bhadra, G.S. Kumar, Interaction of berberine, palmatine, coralyne, and
Anal. Biochem. 352 (2006) 231–242.
sanguinarine to quadruplex DNA: a comparative spectroscopic and
[16] L.S. Lerman, Structural considerations in the interaction of DNA and acridines,
calorimetric study, Biochim. Biophys. Acta 2011 (1810) 485–496.
J. Mol. Biol. 3 (1961) 18–30.
[43] H.A. Benesi, J.H. Hildebrand, Spectrophotometric investigation of the
[17] R. Martínez, L.C. García, The search of DNA-intercalators as antitumoral drugs:
interaction of iodine with aromatic hydrocarbons, J. Am. Chem. Soc. 71
what it worked and what did not work, Curr. Med. Chem. 12 (2005) 127–151.
(1949) 2703–2707.
[18] X. Shui, M.E. Peek, L.A. Lipscomb, Q. Gao, C. Ogata, B.P. Roques, C. Garbay-
[44] J. Liu, T. Zhang, T. Lu, L. Qu, H. Zhou, Q. Zhang, L. Ji, DNA-binding and cleavage
Jaureguiberry, A.P. Wilkinson, L.D. Williams, Effects of cationic charge on
studies of macrocyclic copper(II) complexes, J. Inorg. Biochem. 91 (2002) 269–
three-dimensional structures of intercalative complexes structure of a bis-
276.
intercalated DNA complex solved by MAD phasing, Curr. Med. Chem. 7 (2000)
[45] M. Sirajuddin, S. Ali, A. Haider, N.A. Shah, A. Shah, M.R. Khan, Synthesis,
59–71.
characterization, biological screenings and interaction with calf thymus
[19] M.J. Waring, C. Bailly, The purine 2-amino group as a critical recognition
DNA as well as electrochemical studies of adducts formed by azomethine
element for binding of small molecules to DNA, Gene 149 (1994) 69–79.
[2-((3,5-dimethylphenylimino)methyl)phenol] and organotin(IV) chlorides,
[20] C. Rehn, U. Pindur, Molecular modeling of intercalation complexes of
Polyhedron 40 (1) (2012) 19–31.
antitumor active 9-aminoacridine and a [d, e]-anellated isoquinoline
[46] M. Sirajuddin, S. Ali, N.A. Shah, M.R. Khan, M.N. Tahir, Synthesis,
derivative with base paired deoxytetranucleotides, Monatsh. Chem. 127
characterization, biological screenings and interaction with calf thymus DNA
(1996) 645–658.
of a novel azomethine 3-((3,5-dimethylphenylimino)methyl)benzene-1,2-diol,
[21] M. Baginski, F. Fogolari, J.M. Briggs, Electrostatic and non-electrostatic
Spectrochim. Acta A 94 (2012) 134–142.
contributions to the binding free energies of anthracycline antibiotics to
[47] F. Arjmand, A. Jamsheera, DNA binding studies of new valine derived chiral
DNA, J. Mol. Biol. 274 (1997) 253.
complexes of tin(IV) and zirconium(IV), Spectrochim. Acta A 78 (2011) 45–51.
[22] W. Bauer, J. Vinograd, The interaction of closed circular DNA with intercalative
[48] G. Pratviel, J. Bernadou, B. Meunier, DNA and RNA cleavage by metal
dyes: III. Dependence of the buoyant density upon superhelix density and base
complexes, Adv. Inorg. Chem. 45 (1998) 251.
composition, J. Mol. Biol. 54 (1970) 281–298.
[49] N. Shahabadi, S. Kashanian, M. Khosravi, M. Mahdavi, Multispectroscopic DNA
[23] S. Neidle, Z. Abraham, Structural and sequence-dependent aspects of drug
interaction studies of a water-soluble nickel(II) complex containing different
intercalation into nucleic acid, CRC Crit. Rev. Biochem. 171 (1984) 73–121.
dinitrogen aromatic ligands, Tran. Met. Chem. 35 (2010) 699–705.
[24] M.V. Keck, S.J. Lippard, Unwinding of supercoiled DNA by platinum–ethidium
[50] K.A. Kumar, K.L. Reddy, S. Vidhisha, S. Satyanarayana, Synthesis,
and related complexes, J. Am. Chem. Soc. 114 (1992) 3386–3390.
characterization and DNA binding and photocleavage studies of
[25] C. Moucheron, A.K.D. Mesmaeker, New DNA-binding ruthenium(II) complexes
[Ru(bpy)2BDPPZ]2+, [Ru(dmb)2BDPPZ]2+ and [Ru(phen)2BDPPZ]2+ complexes
as photo-reagents for mononucleotides and DNA, J. Phys. Org. Chem. 11 (1998)
and their antimicrobial activity, Appl. Organometal. Chem. 23 (2009)
577–583.
409–420.
[26] I. Haq, P. Lincoln, D. Suh, B. Norden, B. Chowdhry, J. Chaires, Interaction of
[51] G.N. KaluCerovic, R. Paschkea, S. Prashar, S. Gomez-Ruiz, Synthesis,
DELTA-and LAMBDA-[Ru(phen)2DPPZ]2+ with DNA: a calorimetric and
characterization and biological studies of 1-D polymeric triphenyltin(IV)
equilibrium binding study, J. Am. Chem. Soc. 117 (1995) 4788–4796.
carboxylates, J. Organomet. Chem. 695 (2010) 1883–1890.
[27] P. Lincoln, B. Norden, DNA binding geometries of ruthenium(II) complexes
[52] M.L. Falcioni, M. Pellei, R. Gabbianelli, Interaction of tributyltin(IV) chloride
with 1,10-phenanthroline and 2,20 -bipyridine ligands studied with linear
and a related complex [Bu3Sn(LSM)] with rat leukocytes and erythrocytes:
dichroism spectroscopy. Borderline cases of intercalation, J. Phys. Chem. B 102
effect on DNA and on plasma membrane, Mutat. Res. 653 (2008) 57–62.
(1998) 9583–9594.
M. Sirajuddin et al. / Journal of Photochemistry and Photobiology B: Biology 124 (2013) 1–19 19

[53] S.A. Tysoe, R.J. Morgan, A.D. Baker, T.C. Strekas, Spectroscopic investigation of [65] W.Y. Li, J.G. Xu, X.Q. Guo, Q.Z. Zhu, Y.B. Zhao, Study on the interaction between
differential binding modes of DELTA- and LAMBDA-Ru(bpy)2 (ppz)2+ with calf rivanol and DNA and its application to DNA assay, Spectrochim. Acta A 53 (5)
thymus DNA, J. Phys. Chem. 97 (1993) 1707–1711. (1997) 781–787.
[54] S. Zhang, Y. Zhu, C. Tu, H. Wei, Z. Yang, L. Lin, J. Ding, J. Zhang, Z. Guo, A novel [66] N. Li, Y. Ma, L. Guo, X. Yang, Interaction of anticancer drug mitoxantrone with
cytotoxic ternary copper(II) complex of 1,10-phenanthroline and L-threonine DNA analyzed by electrochemical and spectroscopic methods, Biophys. Chem.
with DNA nuclease activity, J. Inorg. Biochem. 98 (2004) 2099–2116. 116 (2005) 199–205.
[55] S. Dey, S. Sarkar, H. Paul, E. Zangrando, P. Chattopadhyay, Copper(II) complex [67] H.L. Wu, K. Li, T. Sun, F. Kou, F. Jia, J.K. Yuan, B. Liu, B.L. Qi, Synthesis, structure,
with tridentate N donor ligand: synthesis, crystal structure, reactivity and DNA and DNA-binding properties of manganese(II) and zinc(II) complexes with tris
binding study, Polyhedron 29 (2010) 1583–1587. (N-methylbenzimidazol-2-ylmethyl) amine ligand, Tran. Met. Chem. 36
[56] K. Ghosh, P. Kumar, N. Tyagi, U.P. Singh, V. Aggarwal, M.C. Baratto, Synthesis (2011) 21–28.
and reactivity studies on new copper(II) complexes: DNA binding, generation [68] H. Wu, F. Jia, F. Kou, B. Liu, J. Yuan, Y. Bai, A Schiff base ligand N-(2-
of phenoxyl radical, SOD and nuclease activities, Eur. J. Med. Chem. 45 (2010) hydroxylacetophenone)-3-oxapentane-1,5-diamine and its nickel(II) complex:
3770–3779. synthesis, crystal structure, antioxidation, and DNA-binding properties, Tran.
[57] K. Ghosh, P. Kumar, N. Tyagi, Synthesis, crystal structure and DNA interaction Met. Chem. 36 (2011) 847–853.
studies on mononuclear zinc complexes, Inorg. Chim. Acta 375 (1) (2011) 77– [69] S. Anbu, M. Kandaswamy, Electrochemical, magnetic, catalytic, DNA binding
83. and cleavage studies of new mono and binuclear copper(II) complexes,
[58] S. Shujha, A. Shah, Z. Rehman, N. Muhammad, S. Ali, R. Qureshi, N. Khalid, A. Polyhedron 30 (2011) 123–131.
Meetsma, Diorganotin(IV) derivatives of ONO tridentate Schiff base: synthesis, [70] M.T. Carter, A.J. Bard, Voltammetric studies of the interaction of tris(1,10-
crystal structure, in vitro antimicrobial, anti-leishmanial and DNA binding phenanthroline)cobalt(III) with DNA, J. Am. Chem. Soc. 109 (1987) 7528–7530.
studies, Eur. J. Med. Chem. 45 (2010) 2902–2911. [71] Q. Feng, N.Q. Li, Y.Y. Jiang, Electrochemical studies of porphyrin interacting
[59] C.C. Ju, A.G. Zhang, C.L. Yuan, X.L. Zhao, K.Z. Wang, The interesting DNA- with DNA and determination of DNA, Anal. Chim. Acta 344 (1997) 97–104.
binding properties of three novel dinuclear Ru(II) complexes with varied [72] Y.T. Li, W. Sun, Z.Y. Wu, Y.J. Zheng, C.W. Yan, Synthesis, structure and
lengths of flexible bridges, J. Inorg. Biochem. 105 (2011) 435–443. voltammetric studies of copper(II) polymer with DNA interaction: the first 1-D
[60] Z. Rehman, A. Shah, N. Muhammada, S. Ali, R. Qureshi, A. Meetsma, I.S. Butler, coordination polymer alternately bridged by oxalate and oxamidate ligands, J.
Synthesis, spectroscopic characterization, X-ray structures and evaluation of Inorg. Organomet. Polym. 20 (2010) 586–591.
binding parameters of new triorganotin(IV) dithiocarboxylates with DNA, Eur. [73] M. Aslanoglu, N. Oge, Voltammetric, UV absorption and viscometric studies of
J. Med. Chem. 44 (2009) 3986–3993. the interaction of norepinephrine with DNA, Turk. J. Chem. 29 (2005) 477–485.
[61] S.S. Kalanur, U. Katrahalli, J. Seetharamappa, Electrochemical studies and [74] M. Aslanoglu, G. Ayne, Voltammetric studies of the interaction of quinacrine
spectroscopic investigations on the interaction of an anticancer drug with DNA with DNA, Anal. Bioanal. Chem. 380 (2004) 658–663.
and their analytical applications, J. Electroanal. Chem. 636 (2009) 93–100. [75] X. Chu, G.L. Shen, J.H. Jiang, T.F. Kang, B. Xiong, R.Q. Yu, Anal. Chim. Acta 373
[62] A. Shah, M. Zaheer, R. Qureshi, Z. Akhter, M.F. Nazar, Voltammetric and (1998) 29.
spectroscopic investigations of 4-nitrophenylferrocene interacting with DNA, [76] T.W. Welch, H.H. Thorp, J. Phys. Chem. 100 (1996) 13829.
Spectrochim. Acta A 75 (2010) 1082–1087. [77] M.T. Carter, M. Rodriguez, A.J. Bard, J. Am. Chem. Soc. 111 (1989) 8901.
[63] J.R. Lakowicz, Principles of Fluorescence Spectroscopy, third ed., Springer, [78] N. Muhammad, A. Shah, Z. Rehman, S. Shuja, S. Ali, R. Qureshi, A. Meetsma,
2006. M.N. Tahir, Organotin(IV) 4-nitrophenylethanoates: synthesis, structural
[64] D. Suh, J.B. Chaires, Criteria for the mode of binding of DNA binding agents, characteristics and intercalative mode of interaction with DNA, J.
Biorg. Med. Chem. 3 (6) (1995) 723–728. Organomet. Chem. 694 (2009) 3431–3437.

You might also like