Vous êtes sur la page 1sur 12

J Fluoresc (2017) 27:2119–2130

DOI 10.1007/s10895-017-2151-x

ORIGINAL ARTICLE

Analytical Techniques Used to Detect DNA Binding Modes


of Ruthenium(II) Complexes with Extended Phenanthroline Ring
C. Shobha Devi1 · B. Thulasiram2 · S. Satyanarayana3 · Penumaka Nagababu2,4 

Received: 11 April 2017 / Accepted: 31 July 2017 / Published online: 22 August 2017
© Springer Science+Business Media, LLC 2017

Abstract  This review describes the analytical techniques Introduction


used to detect DNA-probes such as Ru(II) complexes with
hetero cyclic imidazo phenanthroline (IP) ligands. Studies Heterocyclic systems are one of the most important bioactive
on drug-DNA interactions are useful biochemical techniques molecules found in nature. The compounds containing mainly
for visualization of DNA both in vitro and in vivo. The inter- nitrogen, oxygen and sulphur atom constitute a large class of
actions of small molecules that binds to DNA are mainly compounds used in biological and medicinal applications [1].
classified into two major classes, one involving covalent An enormous number of heterocyclic systems that include
binding and another non-covalent binding. Covalent bind- mainly 6- and 5- membered ring compounds represent a dif-
ing in DNA can be irreversible and may leads to inhibition ferent group of molecular support. Several such heterocyclic
of all DNA processes which subsequently leads to cell death. skeletons have been magnificently incorporated into many novel
Usually, covalent interactions leads to permanent changes in drug leads and therapeutic agents [2–4]. Examples that illustrate
the structure of nucleic acids. The non-covalent interaction the biological importance and therapeutic utility of some het-
of molecules with DNA can be due to electrostatic interac- erocyclic compounds and its substitutions includes improved
tion, intercalation and groove binding. These interactions of 1, 10, phenanathroline compounds like imidazo phenanthroline
DNA probes can be explored by various spectroscopic tech- (IP) ligands [5]. Bioinorganic chemistry (or) Medicinal inor-
niques viz. UV–visible, emission, emission quenching spec- ganic chemistry is rapidly increasing in significance in both
troscopy, viscosity and thermal denaturation measurements. therapeutic and diagnostic medicine [6]. In the last few decades,
ruthenium(II) polypyridyl complexes with imidazo phenanathro-
Keywords  DNA-binding · Intercalation · Emission line (IP) ligands have been developed as DNA structural probes
studies · Ruthenium(II) complexes · Analytical techniques (Scheme 1) [7]. In this review we are describing octahedral metal
complexes containing an intercalating ligand (IP) which binds to
DNA three dimensionally. However, in-order to tune the DNA
C.S. Devi and Penumaka Nagababu contributed equally.
binding ability ­(Kb) of the complexes, its ancillary ligands can
* Penumaka Nagababu be modified or functionalized. A ligand with a small aromatic
babupenumaka@gmail.com or hetero aromatic ring can bind to the DNA in the double helix
1
through intercalating mode, thereby distorting the DNA back-
Department of Chemistry, RGUKT, Basar, Telangana State,
India
bone conformation and interfering the interactions of DNA–pro-
2
tein. Ruthenium(II) complexes are most effective sensitizers due
Inorganic & Physical Chemistry Division, CSIR-Indian
Institute of Chemical Technology, Uppal Road, Tarnaka,
to their strong charge transfer absorption in the visible region.
Hyderabad 500007, Telangana State, India Due to the interesting photo physical properties, chemical stabil-
3
Department of Chemistry, Osmania University, Tarnaka,
ity and biological properties of ruthenium complexes, much of
Hyderabad, Telangana State, India research has been focused on these complexes. More recently,
4
CSIR-NEERI Zonal Laboratory, I‑8, Sector C, East Kolkata,
attention has been paid to the use of these complexes as chemo-
Area Development Project, P.O. East Kolkata, Township, therapeutic agents for cancer [8], DNA conformation probes and
Kolkata 700 107, India DNA cleavage agents [9–12]. When the chemical compound

13
Vol.:(0123456789)

2120 J Fluoresc (2017) 27:2119–2130

Scheme 1  Structures of imidazo phenanthroline (IP) ligands as well as metal complexes

interacts with DNA, some modifications are experienced in mol- and the various analytical methods that can be used to determine
ecule as well as in DNA. Detection and explanation of these intercalative binding mode of ruthenium(II) complexes having
modifications make a great challenge for analytical techniques. IP based ligands.
The aim of the modern research is to understand the specific
binding mode of drug-DNA complex and to find drug binding Deoxyribonucleic Acid (DNA)
sites, as well as conformational changes observed due to the
drug-DNA interaction. In this mini-review, we are mainly focus- DNA (deoxyribonucleic acid) is made up of molecules
ing on the current advancements in the field of bio-analytical called nucleotides and an important complex molecule
methods using ruthenium polypyridine complexes with extended that carries the genetic code. Each nucleotide holds a
phenantroline ring system. Our main focus is how ruthenium(II) sugar group, a phosphate group, and a nitrogen base like;
polypyridyl complexes are being used as DNA targeting agents adenine (A), thymine (T), guanine (G) and cytosine (C).

13
J Fluoresc (2017) 27:2119–2130 2121

The sides of the ladder as strands of alternating phosphate Covalent Binding


and sugar groups-strands those run in conflicting direc-
tions. Each “String” of the ladder is made up of two base Covalent Transition metal complexes can bind covalently to
pairs coupled together by hydrogen bonds due to its spe- DNA. They can occur within each linear strand and strongly
cific nature of chemical pairing, adenine always couple bond the sugars, bases, and phosphate groups [14]. It is
with thymine, and cytosine always couple with guanine irreversibly bound DNA that leads to complete inhibition
(Fig. 1). Within this composition, one ladder mirrors to the of DNA processes and subsequent cell death. Chloraum-
other as a result of the anti-aligned direction of the sugar- bucil is a bi-functional alkylating agent of DNA [15] and
phosphate backbones and also the complementary nature pyrrolo-(1,4)-benzodiazepines interacts with DNA via the
of the adenine–thymine and cytosine-guanine base pair- covalent bond formation [16]. One among the early exam-
ing. The distances between the any two nitrogenous base ples of drugs that interacted covalently with DNA is the
pairs are nearly equal, so that the distance across the DNA well-known anticancer agent, Cisplatin [cis-diammine
double helix is similar. DNA double helix that can bind dichloro platinum (II)]. The inhibition of cell replication is
with the small molecules having negatively charged phos- brought out by the coordination of DNA on N7 and O6 sites
phate backbone is called electrostatic binding, The sites of guanine (Fig. 2) by means of the two labile chlorides of
present in the major and minor grooves which are with the molecule. The distance between the two chloride atoms
hydrogen donating and accepting groups is called hydro- (3.4 ­Ao) is the key factor in the coordination of this drug
gen bonding, the aromatic hydrophobic components can with DNA.
interact with the phosphate oxygen atoms is called van-
der Waals interactions. In generally the cations molecules Non‑covalent Interaction
can interact with electron-rich sites on DNA double helix
such as N-donor ligands is called covalent binding and The non-covalent mode of binding of small molecules-
guanine has the highest electron density of all other four DNA can be classified into three major categories
nitrogenous base pairs. It is accepted that this interaction (Scheme 2):
is answerable of cisplatin effect on cancer cells [9–12].
a) Electrostatic interaction: Due to DNA double helix con-
figuration, negatively charged phosphate groups that are
Mode of Drug‑DNA Interactions found in the periphery, attract metal cation groups by
coulombic forces. These types of interactions are called
Studies on Drug-DNA interactions are useful biochemical electrostatic interactions.
tools for visualization and analysis of molecular networks b) Intercalation: Intercalation occurs when a planar aro-
of DNA both inside and outside the cell [13]. Typically the matic chromophore glide between two adjacent DNA
interactions of small molecules to DNA are characterized base pairs [17]. There are many proved metal polyp-
into two major classes. iridyl complexes are there to bind intercalative with

Fig. 1  Structure of double
stranded DNA exists of two
nucleotide chains whose
nitrogen base pairs are linked by
hydrogen bonds. (Copied from
http://www.SciLinks.org, web
code: cbn-4121)

13

2122 J Fluoresc (2017) 27:2119–2130

Fig. 2  Structures of Guanine
(a), Cisplatin (b) and covalent
binding of Cisplatin (c)

DNA base pairs, which results in increasing in length DNA binding. The intense fluorescence of these agents
of DNA. DNA intercalators are widely used in anti- upon DNA binding derives from electronic perturbations
neoplastic, antitumor, antibiotic, antimalarial and in the ligand and result from the release of fluorescence
antifungal agents, however, that does not mean that all quenching water as well as the stabilization of overlap-
intercalators have genotoxicity (defined by the abil- ping π-systems. The application of fluorescent intercala-
ity to alter the genetic material) in cells as a means of tors to examine the DNA binding characteristics of other
inducing a toxic effect [17]. Intercalation changes the molecules is particularly useful [18]. In addition, ethidium
base pair spacing from 3.4 to 6.8 Å and persuade local bromide is employed to visualize the presence or the loca-
structural changes to the DNA such as helix unwind- tion of DNA fragments in agarose gel electrophoresis [19].
ing and B-form to A-form transitions. These changes
can alter DNA-based processes. Intercalators such as c) Groove binding: There are mainly two grooves in the
ethidium bromide and thiazole orange are known potent DNA double helix, which are called major groove and
mutagens. Similarly, due to their tight association with minor groove. The minor groove occurs where they
DNA, many intercalators have clinical efficacy and are close together and the major groove occurs where
have been used as chemotherapeutic treatments for the backbones are far apart. Generally groove binding
DNA replication in growing cancer cells. drugs like distamycin, Hoescht 33258 are most often
favour to bind the minor groove of DNA double helix
Several intercalators such as ethidium bromide and [20]. As we know that all intercalative or groove bind-
thiazole orange (Fig. 3) are used as non-specific inter- ing drugs can be used for treatment of cancer, tumor
calating agents that display enhanced fluorescence upon and infectious diseases induced by microbial organisms

Scheme 2  Schematic rep-
resentation showing various
techniques to investigate DNA
binding of metal complexes

13
J Fluoresc (2017) 27:2119–2130 2123

IP ligands in metal complexes present in UV region i.e.


below 400 nm as shown in Fig. 4 (Table 1).
Bathochromism (red shift) as well as hypochromism were
observed when a polypyridyl complex with planar aromatic
ligands bound to DNA through intercalation mode and
strong π-π* stacking interaction between base pairs of the
DNA and the planar aromatic ligands occurred [34]. It is
generally agreed that the extent of the hypochromic shift in
Fig. 3  Structures of ethidium bromide (1) and thiazole orange (2) the UV–Visible spectrum is consistent with the resistance of
(both are proven Intercalators)
intercalative interaction [35]. The mixture of ruthenium and
DNA solutions were allowed to stand for 2 to 3 min before
[21]. For instance, the anticancer drug mitomycin is recording the spectra. Same titrations should be repeated
a typical groove binder and a class of clinically most until there was no change is observed in the absorption
important compound anthracyclines with anti-bacterial spectrum. In order to compare quantitatively the binding
and anti-neoplastic properties, take advantage of both strength ­(Kb) of the Ru-DNA were calculated by observ-
modes of binding as a groove-binding as well as an ing the changes in spectra at their respective metal–ligand
intercalative of DNA [22]. charge transfer (MLCT) with increasing DNA stock solu-
tion. Generally the following equation is used to calculate
the ­Kb [36].

DNA Binding Studies of Metal Complexes


/( ) /( ) /( ( ))
[DNA] εa − εf = [DNA] εb − εf + 1 Kb εb − εf ,
with Imidazo Phenanthroline (IP) Ligands
On plotting [DNA]/(Ɛa–Ɛf) vs [DNA], the binding con-
Over the past few years significant progress has been made stant ­(Kb) is obtained by the ratio of the slope to the intercept
to evaluate transition metal complexes as structural probes (Fig. 4). The concentration of DNA will be given as [DNA],
of DNA. The number of analytical techniques developed to Ɛa, Ɛf and Ɛb corresponds to the apparent absorption coef-
evidence metal complex-DNA interactions. Various physi- ficient ­Aobsd/[Ru complex], the extinction coefficient for the
cal, biochemical methodologies are employed to detect the free complex, and the extinction coefficient for the complex
DNA interactions [23–26]. The classical UV–vis spectro- in the fully bound form respectively.
photometry, fluorescence spectroscopy, viscosity and/or The Fig.  4 shows the absor ption spectra of
thermal melting temperature studies have been extensively [Ru(phen)2ptip]2+ complex in the presence and absence of
reported for the detection of DNA interactions of different DNA stock. Complex alone in aqueous solution showed
metal complexes. very low-energy MLCT absorption at wavelength 470 and
620 nm. On increasing the DNA, spectral changes showed
Absorption Spectroscopy great hypochromism (Fig. 4) which suggested that the com-
plex was strongly bound to DNA, due to stacking interaction
Absorption spectroscopy method is one of the much useful between the aromatic intercalative ligand (IP ligand) and the
for investigation on interplay of ruthenium (II) complexes DNA base pairs.
with CT-DNA. By utilizing absorption spectroscopy tech-
nique in DNA binding analysis with ruthenium (II) com- Emission Spectroscopy
plexes has been studied by various research groups [27, 28].
Generally transition metal ions are coloured in solutions due Fluorescence spectroscopy is most useful method to
d-orbitals of ions split in the higher and lower energy lev- study the ruthenium(II) complex-DNA binding interac-
els, when d-orbitals of the ruthenium (II) complexes interact tions. Complexes can strongly bind with DNA and be
with the negatively charged complexing agents (ligands) in guarded by DNA efficiently, since the hydrophobic envi-
such a way they become non-degenerate. The colour of solu- ronment inside the DNA helix reduces the intelligibility
tions is affected by certain ligands which are coordinated of solvent water molecules to the complex and their by
with coordination centre. For instance, imidazophenanthro- restricting the mobility of the complex at the binding site,
line (IP) ligands, due to their extended conjugation increase which results in increasing the intensity of fluorescence.
the absorption of metal complexes in visible region. In Emission quenching experiments of Ru(II) complexes are
Ru(II) polypyridyl complexes a metal and a ligand permit to usually performed using benchmark quencher like potas-
charge transfer band which is observed in the region between sium iodide and ferrocyanide to detect the binding of the
420 and 490 nm [8–12, 29–33]. The π-π* transitions of the complexes with DNA. Further, the quenching constant

13

2124 J Fluoresc (2017) 27:2119–2130

Fig. 4  Absorption spectra of
complexes [Ru(phen)2ptip]2+ in
Tris–HCl buffer upon addition
of CT-DNA. Arrow shows the
hypsochromic and bathochro-
mic shift upon increase the
DNA concentration. Plots of
[DNA]/ (εa-εf) vs [DNA] for the
titration of DNA with Ru(II)
complexes

­(Ksv) can be calculated by using Stern–Volmer relation- binding constants are calculated using the equation [38]
ship [37]. Ruthenium complexes with polypyridyl ligands and substitution the data into the Scatchard plot of r vs r/
also possess a fascinating phenomenon called the light- Cf. Where ‘r’ is the binding ratio C
­ b/[DNA] and C­ f is the
switch effect. When the polypyridyl ligand (IP) is inter- free complex. The affinity of binding for this complex
calated to DNA, embedded in lipid membranes or in an was reported as 5.02 × 10 4 ­M − 1. The emission binding
organic environment, the complex is strongly luminescent constants for various Ru(II) metal complexes is shown
but in aqueous solution the emission is totally quenched in Table 1.
[29–31].
This quality of the complexes makes them interest- Fluorescence Quenching
ing probes to study cells by using fluorescence micros-
copy. Emission spectral experiments were conducted at Decrease in fluorescence intensity is called fluores-
ambient conditions to find the binding constant between cence quenching. It is very useful technique to evalu-
the ruthenium complexes and DNA. Emission titrations ate DNA binding affinity if a fluorophore (fluorescent
were performed similar to the absorption titrations. To chemical compound that can re-emit light upon light
the fixed concentrations of Ru(II) complex aliquots of excitation) undergoes a change in an environment in the
DNA were added until there is no change in the intensity binding process that is reflected by change intensity in
of spectrum. As shown in Fig. 5 with increase in concen- fluorescence. The molecular interactions between metal
tration of DNA, the intensities of [Ru(en) 2pyip]2+ com- complexes and DNA can lead to reduction of the fluo-
plex increased by around 2–10 times. The fluorescence rescence intensity of fluorophore. This experiment is

Table 1  Results of absorption Complexes Absorbance λmax (nm) Kb ­(M− 1) from Kb ­(M− 1) from emission
and emission binding constants absorbance and references
­(Kb) for Ru(II) complexes
[Ru(phen)2(PtIP)]2+ 452,350, 275 7.01 × 105 6.10 × 105 [30]
[Ru(phen)2(IPPIP)]2+ 298, 366, 446 2.1 ± 0.2 × 105 [31]
[Ru(bpy)2(IPPIP)]2+ 256, 288, 446 8.5 ± 0.2 × 104 [31]
[Ru(bpy)2(PtIP)]2+ 451,342, 280 3.87 × 105 4.02 × 105
[30, 31]
[Ru(en)2(AIP)]2+ 437, 212 2.5 × 105 [32]
[Ru(en)2(PYIP)]2+ 452, 257 1.3 × 105 [32]
[Ru(en)2IP]2+ 277, 457 4.8 ± 0.2 × 104 [33]
[Ru(en)2PIP]2+ 257, 467 5.7 ± 0.1 × 104 [33]

13
J Fluoresc (2017) 27:2119–2130 2125

Fig. 5  Emission spectra of
[Ru(en)2pyip]2+ [32, 33] with
addition of DNA, well separated
band at 600 nm Scatchard plot
was drawn with r/Cf vs r, where
r is the ­Cb/[DNA] and C
­ f is the
concentration of free complex

usually performed using [Fe(CN)6]4− as quencher which DNA Photocleavage


further supports the above proposal in emission experi-
ment. Stern–Volmer plots of various complexes reported Electrophoresis through agarose is the standard method
indicates that ruthenium(II) complex alone (absence used to separate, identify or purify DNA fragments [26,
of DNA) efficiently quenched by [Fe(CN) 6] 4−. On the 42]. The technique is simple, rapid to perform and capable
other hand solution with complex + DNA (presence of resolving fragments of DNA that cannot be separated by
of DNA) quenching was less, because [Fe(CN) 6 ] 4− is other procedures. In general, electrophoretic separation is
highly negatively charged molecule it can repelled less carried out on agarose gels for DNA samples. Using this
negatively charged DNA phosphate backbone and would technique, bands containing as little 1–50 μg of DNA can
inhibit the quenching of the emission intensity of the also be detected by direct examination of the agarose gel
DNA bound complex. This may also explain by the fac-
tor that the bound metal cations of the ruthenium(II)
complexes protected from the anionic water-bound
[Fe(CN) 6 ] 4− (quencher) by DNA (negatively charged)
phosphate backbone hindering decreasing the emission
intensity of ruthenium(II) complexes [39, 40]. By using
Stern–Volmer equation the emission quenching constant
­K sv can be calculated [41]. The binding affinity can be
taken from measured slop.

I0 ∕I = 1 + Ksv [Q]

where I and I­ 0 are the fluorescence intensities in the


presence and absence of [Fe(CN)6]4− (quencher) respec-
tively, Q is the concentration of the [Fe(CN) 6] 4−, ­K sv is
a linear Stern–Volmer constant. In the quenching plot
of ­I 0/I versus [Q], K
­ sv is given by the slope. The ferro-
cyanide quenching plot for complex [Ru(phen) 2ptip] 2+
in the absence, presence and excess of DNA is shown
in Fig. 6 complex has shown linear Stern–Volmer plots. Fig. 6  Emission quenching of [Ru(phen)2ptip]2+ complex with
This confirms that the complex can bind to DNA more ­ 4[Fe(CN)6] in the absence (a) and presence (b) [Ru] = 20  mM and
K
strongly. excess of DNA (c)

13

2126 J Fluoresc (2017) 27:2119–2130

in the UV light [43, 44]. When an electric field is applied for solutions of DNA with and without Ru(II) complex as a
across the gel, DNA, which is negatively charged at neu- function of temperature.
tral pH will migrates toward the anode. The interaction of The thermal data are plotted as (A–A0)/(Af–A0) versus
pBR322 DNA with ruthenium(II) polypyridyl complexes thermal denaturation temperature, where A, A ­ 0, and ­Af are
was studied in order to determine the efficiencies of sensi- the observed, initial, and the final absorbance at 260 nm,
tize DNA cleavage. This objective was achieved by moni- respectively. This experiment can also be carried out with
toring the transition from the naturally occurring, cova- DNA in the absence of metal complexes which reveals
lently closed circular form (Form I) to the nicked circular that the Tm for the duplex is 60.6 ± 0.5 °C under same
relaxed form (Form II) by means of gel electrophoresis of experimental conditions. As shown in Fig. 8 the observed
the plasmid. The cleavage of supercoiled DNA (Form I) change melting temperatures (∆T m) in the presence of
to the nicked circular DNA (Form II) was observed for all ruthenium(II) complexes including [Ru(phen) 2pyip] 2+
the complexes regardless of different incubation periods and [Ru(phen) 2aip] 2+ are ∼ 5–10  °C, respectively. The
and different concentrations of the complexes. The results large change in T ­ m suggests that the intercalative binding
of the experiments carried out in the concentration range affinities of ruthenium (II) complexes with DNA were very
10–40 µM for all the three complexes after 1 and 2 h of strong [47].
incubation. This study demonstrates that there is no sig-
nificant cleavage for controls. All complexes cleave DNA Viscosity Measurements
and increase in concentration of the complex increases the
cleavage of the DNA (Fig. 7). Viscosity is a measure of a fluid’s resistance to flow. It
defines the internal friction of a moving liquid. A liq-
uid with little viscosity flows easily because its molecu-
To Know Mode of Binding with DNA lar makeup results in very slight friction when it is in
motion. A liquid with large viscosity resists motion since
Thermal Denaturation its molecular makeup gives it a lot of internal friction.
In this study the viscosity measurements were carried
The DNA mode of binding to metal complexes can moni- out for further clarification of the interaction between
tor with the thermal denaturation ­(Tm) method. When the the ruthenium complexes with DNA. As per Kelly et al.,
temperature of the (DNA + complex) solution increases if increase of DNA length in between base pairs accom-
the double stranded DNA relatively separate in to single modate at intercalation locations, hence it will increase
strands, and intensity of absorption spectra is increases of the viscosity of DNA solution, which are properties of
the DNA bases at around 260 nm [20]. As per previous lit- intercalation binding mode of complexes. By contrast,
erature reports [45, 46] the intercalation binding of a mol- complexes that binds groove modes in the DNA by par-
ecule (metal complexes) usually results in a considerable tial/or non-classical intercalation typically cause less
increase of ­Tm. The duplex DNA structure can stabilize or no change in DNA viscosity [48–50]. From the lit-
when the metal complexes which can intercalate between the erature it is evident that with proven intercalative DNA
base pairs of DNA. The hyperchromic effect in the absorp- binder ethidium bromide under appropriate conditions
tion spectra of the DNA bases at 260 nm are recorded in causes a significant increase in the DNA length. In con-
the temperature range 40 to 80 °C. The denaturation curves trast, a partial intercalation ligand could bend the DNA
are acquired by computing the absorbance at λmax 260 nm helix, decreases its length and concurrently its viscosity.

Fig. 7  Photocleavage studies of
pBR322 DNA, in the absence
and presence of ruthenium(II)
complexes {[Ru(phen)2ptip]2+,
[Ru(bpy)2ptip]2+ and
[Ru(dmb)2ptip]2+} light after
30 min irradiation at 365 nm.
Lane 0 control untreated
plasmid DNA three sets of 4
lanes with 20, 40, 60 and 80 µM
respectively

13
J Fluoresc (2017) 27:2119–2130 2127

5 viscosity of DNA alone [5, 51]. A ligand that binds in


a the DNA groove causes a less pronounced change or no
b change in the viscosity of a DNA solution. The effect in
4 c presence of relative viscosity are calculated for a com-
plex to DNA ratio between 0.0 and 0.15 adopting the
d
method described by Cohen and Eisenberg [52]. Vis-
Relative OD

3 e cosity measurements can be carried out on calf thymus


DNA by varying the concentration of the added com-
plex. Effect of increasing amount of ethidium bromide
2
(A), Δ-[Ru(phen)3]2+ (E), Λ-[Ru(phen) 3]2+ (B), and rac-
[Ru(phen) 3] 2+ (D) and Hoechst 33,258 (C) on relative
viscosity of CT-DNA are shown in Fig. 9.
1
40 45 50 55 60 65 70 75 80 Ethidium Bromide Displacement Studies
Temperature 0C

Since 1950s the Ethidium bromide (EtBr) is used in medi-


Fig. 8  Plots of Relative O.D vs temperature of CT DNA (90  µm) cine to treat trypanosomiasis in veterinary for cattle a dis-
and complexes, DNA alone (a), DNA + complex [Ru(phen)2pyip]2+
(b), DNA + complex [Ru(phen)2aip]2+ (c), DNA + complex ease caused by trypanosomes [53]. It is very commonly used
[Ru(bpy)2pyip]2+ (d), and DNA + complex [Ru(bpy)2aip]2+ (e) as nucleic acid fluorescent tag in various laboratory tech-
niques. Due to its unique structure, it can easily intercalate
into DNA. The reason for its potent fluorescence after bind-
Viscosity experiments data will be presented as (η/ηo)1/3 ing with DNA is the hydrophobic environment found in the
vs the concentration of [Ru(II)]/[DNA] where η is the DNA base pairs [54]. Ethidium Bromide (EtBr) is a usually
viscosity of DNA in presence of complexes and ηo is the used dye for DNA and RNA detection. It has characteristic

Fig.  9  a Effect of increasing amount of proven intercalator eth- (d). b Effect of increasing amount of complexes on the relative vis-
idium bromide (a), proven groove binder Hoechst 33,258 (c) and for cosities of CT-DNA increases at room temperature and models for
Δ-[Ru(phen)3]2+ (e), Λ-[Ru (phen)3]2+ (b), and rac-[Ru(phen)3]2+ intercalative binding of complexes with DNA at room temperature

13

2128 J Fluoresc (2017) 27:2119–2130

polypyridyl complexes which were reported previously as


intercalators. Viscosity studies (Fig. 9) reveals that EtBr
as intercalator and Δ-[Ru(phen)3]2+, Λ-[Ru(phen)3]2+, and
rac-[Ru(phen)3]2+ are groove binders. However, when we
studied the extended ring system of phenanthroline such
as imidazo-phenanthroline(IP) ligands, they showed good
intercalation with DNA. This suggests that there is a bound-
ary between 1, 10-phenanthroline and imidazo-phenathro-
line (IP) ligands. Imidazo-phenathroline ligands having
moiety’s like phenyl called phenyl-imidazo-phenanthroline
(PIP) and substituted phenyls or extended PIP ligands (like:
AIP, PYIP and PTIP) are good intercalators as well as strong
DNA binders. It was further examined by DNA fragmenta-
tion assay to observe where DNA is cleaved into oligonu-
cleosomal fragments.
AIP, 2-(9-Anthryl)-1H-imidazo[4,5-f][1,10] phenanthro-
line; bpy, 2, 2′, bipyridine; en, ethylenediamine; CT DNA,
Fig. 10  Fluorescence intensity change on addition of ruthenium(II) Calf thymus DNA; EtBr, Ethidium bromide; IP, Imidazo-
complexes (from Ref. [55]) in increasing amounts (0.8  µM in each
phenantrholine; phen, 1, 10, Phenanathroline; PIP, [2-phe-
addition from 1 to 17  µM) to EtBr (9.1  µM)-CT-DNA (6.48  µM)
complex solution nylimidazo [4,5-f][1,10] phenanthroline]; PPIP, [2-(4′-phe-
noxy-phenyl) imidazo[4,5-f][1,10] phenanthroline]; ptip,
(2-(5-phenylthiophen-2-yl)-1H-imidazo[4,5f][1,10] phen-
UV absorbance maxima at 300 and 360 nm, in addition, anthroline; PyIP, 2-(1-pyrenyl)-1H-imidazo[4,5-f][1,10]
observation of the new excited peak at 260 nm radiation phenanthroline
when its reaction with nucleotides and re-emit at 590 nm
as yellow/orange light. Another useful strategy for compete Acknowledgements  This work was suppor ted by CSIR-
NEERI/KZC and IICT Hyderabad, KRC No.: CSIR-NEERI/
analyse was carried out for all complexes with constant of KRC/2017/MAR/KZL/1.
DNA and EtBr concentrations with increasing ruthenium
complexes concentrations. As the complex concentration
increases the fluorescence of EtBr decreased (Fig. 10). This
may be ruthenium complexes compete with EtBr for the References
DNA-binding sites thus displace the EtBr whose intensity
of fluorescence is increased upon DNA intercalative bind- 1. Young DW (1975) Heterocyclic chemistry. Longman, London
ing or it might be quenching on DNA. The results confirm 2. Kumar A, Kumar R (2011) A review on synthesis of Schiff’s bases
of 2-amino 4-phenyl thiazole. Int Res J Pharm 2: 11–12
that all the ruthenium complexes bind robustly in intercala-
3. Patel NB, Shaikh FM (2010) New 4-thiazolidinones of nicotinic
tive mode to DNA. The UV–vis spectral study on DNA- acid with 2-amino-6-methylbenzothiazole and their biological
intercalative binding of complexes to displacement of EtBr activity. Sci Pharm 78: 753–66
from EtBr + DNA solution indicative of intercalative in 4. Bala S, Kamboj S, Kumar A (2010) Heterocyclic 1, 3, 4-oxadia-
zole compounds with diverse biological activities: a comprehen-
nature. However, the extensively studied DNA intercalative
sive review. J Pharm Res 3:2993–2997
agent, ethidium bromide has been used as a probe to identify 5. Nagababu P, Barui AK, Thulasiram B, Devi CS, Satyanarayana
the unknown intercalative agent by displacement of bound S, Patra CR, Sreedhar B (2015) Antiangiogenic activity of mono-
ethidium bromide. Results of ethidium bromide displace- nuclear copper (II) polypyridyl complexes for the treatment of
cancers. J Med Chem 58:5226–5241
ment from DNA + ethidium bromide complex solution by
6. Storr T, Thompson KH, Orvig C (2006) Design of target-
ruthenium(II) complexes illustrates the strong quenching ing ligands in medicinal inorganic chemistry. Chem Soc Rev
activity. 35:534–544
7. Incesu Z, Bljnkli K, Akalin G, Gundogdukaraburu N (2013) The
effects of some phenanthroline ruthenium (Ii) complexes on A549
cell proliferation. Turk J Pharm Sci 10:193–203
Conclusions 8. Devi CS, Satyanarayana S (2012) Review: synthesis, characteri-
zation, and DNA-binding properties of Ru (II) molecular “light
In this review we have discussed in detail about the tech- switch” complexes. J Coord Chem 65:474–486
9. Devi CS, Nagababu P, Natarajan S, Deepika N, Reddy PV,
niques which are useful to determine DNA-binding modes Veerababu N, Singh SS, Satyanarayana S (2014) Cellular uptake,
of ruthenium (II) complexes with modified 1,10 phen- cytotoxicity, apoptosis and DNA-binding investigations of Ru (II)
anthroline lignads. We illustrated a few ruthenium (II) complexes. Eur J Med Chem 72:160–169

13
J Fluoresc (2017) 27:2119–2130 2129

10. Berners-Price SJ, Appleton TG, Kelland LR (2000) The chemistry of 32. Nagababu PE, Shilpa MY, Mustafa MD, Ramjee P, Satyanaray-
Cisplatin in aqueous solution. In: Kelland LR, Farrell NP (eds) Plati- ana S (2008) DNA-binding and photocleavage studies of ethyl-
num-based drugs in cancer therapy. Humana Press, Totowa, pp 3–35 enediamine cobalt (III) and ruthenium (II) mixed ligand com-
11. Chaires JB (1998) Drug-DNA interactions. Curr Opin Struct Biol plexes. Inorg React Mech 6: 301–11
8:314–320 33. Nagababu P, Latha J, Satyanarayana S (2006) DNA-binding
12. Jakupec MA, Galanski M, Keppler BK (2003) Tumour-inhibiting studies of mixed-ligand (Ethylenediamine) ruthenium (II) com-
platinum complexes-state of the art and future perspectives. Rev plexes. Chem Biodivers 3:1219–1229
Physiol Biochem Pharmacol 146:1–53 34. Tan LF, Chao H, Li H, Liu YJ, sun B, Wei W, Ji LN (2005)
13. Schmidt CE, Möller J, Hesslau U, Bauer M, Gabbert T, Kremer B Synthesis, characterization, DNA-binding and photocleavage
(2005) Universitätskliniken im Spannungsfeld des Krankenhaus- studies of [Ru (bpy) 2 (PPIP)] 2 + and [Ru (phen) 2 (PPIP)] 2+.
marktes. Anaesthesist 54:694–702 J Inorg Biochem 99:513–520
14. Thulasiram B, Kumar YP, Aerva RR, Satyanarayana S, Nagababu 35. Pyle AM, Rehmann JP, Meshoyrer R, Kumar CV, Turro NJ,
P (2017) Correlation between molecular modelling and spectro- Barton JK (1989) Trans-dichlorobis(N-p-tolylpyridin-2-amine)
scopic techniques in investigation with DNA binding interaction palladium(II): synthesis, structure, fluorescence features and
of ruthenium (ii) complexes. J Fluoresc 27:587–594 DNA binding. J Am Chem Soc 111:3051–3058
15. Mattes WB, Hartley JA, Kohn KW (1986) DNA sequence selec- 36. Wolfe A, Shimer GH Jr, Meehan T (1987) Polycyclic aromatic
tivity of guanine–N7 alkylation by nitrogen mustards. Nucleic hydrocarbons physically intercalate into duplex regions of dena-
Acid Res 14:2971–2987 tured DNA. BioChemistry 26:6392–6396
16. Hurley LH, Petrusek R (1979) Proposed structure of the anthramy- 37. Joseph R, Lakowicz GW (1973) Quenching of fluorescence by
cin–DNA adduct. Nature 282:529–531 oxygen. Probe for structural fluctuations in macromolecules.
17. Lerman LS (1961) Structural considerations in the interaction of BioChemistry 12:4161–4170
DNA and acridines. J Mol Biol 3:18IN13–30IN14 38. McGhee JD, von Hippel PH (1974) Theoretical aspects of DNA-
18. Boger DL, Winston CT (2001) Thiazole orange as the fluores- protein interactions: co-operative and non-co-operative binding
cent intercalator in a high resolution fid assay for determining of large ligands to a one-dimensional homogeneous lattice. J
DNA binding affinity and sequence selectivity of small molecules. Mol Biol 86:469–489
Bioorg Med Chem 9:2511–2518 39. Kumar CV, Barton JK, Turro NJ (1985) Photophysics of ruthe-
19. Tse WC, Boger DL (2004) A fluorescent intercalator displacement nium complexes bound to double helical DNA. J Am Chem Soc
assay for establishing DNA binding selectivity and affinity. Acc 107:5518–5523
Chem Res 37:61–69 40. Barton JK, Goldberg JM, Kumar CV, Turro NJ (1986) Tris
20. Waring MJ (1965) Complex formation between ethidium bromide (phenanthroline) ruthenium (II) enantiomers with nucleic. J
and nucleic acids. J Mol Biol 13:269–282 Am Chem Soc 108:2081–2088
21. Crawford LV, Waring MJ (1967) Supercoiling of polyoma virus 41. Ghosh BK, Chakravorty A (1989) Electrochemical studies of
DNA measured by its interaction with ethidium bromide. J Mol ruthenium compounds part I. Ligand oxidation levels. Coord
Biol 25:23–30 Chem Rev 95:239–294
22. Takenaka S, Takagi M (1999) Threading intercalators as a new 42. Fisher MP, Dingman CW (1971) Role of molecular conforma-
DNA structural probe. Bull Chem Soc Jpn 72:327–337 tion in determining the electrophoretic properties of polynu-
23. Wilson K, Walker J (2010) Principles and techniques of bio- cleotides in agarose-acrylamide composite gels. BioChemistry
chemistry and molecular biology. Cambridge University Press, 10:1895–1899
Cambridge 43. Aaij C, Borst P (1972) The gel electrophoresis of DNA. Biochim
24. Spiro TG (1980) Nucleic acid-metal ion interactions. Krieger Pub Co Biophys Acta 269:192–200
25. Long EC, Barton JK (1990) On demonstrating DNA intercalation. 44. Sharp PA, Sugden B, Sambrook J (1973) Detection of two restric-
Acc Chem Res 23:271–273 tion endonuclease activities in Haemophilus parainfluenzae using
26. Joseph S, David WR (2001) Molecular cloning: a laboratory analytical agarose-ethidium bromide electrophoresis. BioChem-
manual. Gold Spring Harbor, New York istry 12:3055–3063
27. Morgan RJ, Chatterjee S, Baker AD, Strekas TC (1991) Effects 45. Neyhart GA, Grover N, Smith SR, Kalsbeck WA, Fairley TA, Cory
of ligand planarity and peripheral charge on intercalative binding M, Thorp HH (1993) Binding and kinetics studies of oxidation of
of Ru (2, 2′-bipyridine) 2L2 + to calf thymus DNA. Inorg Chem DNA by oxoruthenium (IV). J Am Chem Soc 115:4423–4428
30:2687–2692 46. Liu YJ, Chao H, Tan LF, Yuan YX, Wei W, Ji LN (2005) Interac-
28. Tysoe SA, Morgan RJ, Baker AD, Strekas TC (1993) Spectro- tion of polypyridyl ruthenium (II) complex containing asymmetric
scopic investigation of differential binding modes of ∆-and ligand with DNA. J Inorg Biochem 99:530–537
Λ-Ru (bpy) 2 (ppz) 2 + with calf thymus DNA. J Phy Chem 47. McGhee JD (1976) Theoretical calculations of the helix–coil
97:1707–1711 transition of DNA in the presence of large, cooperatively binding
29. Devi CS, Nagababu P, Shilpa M, Kumar YP, Reddy MR, Gabra ligands. Biopolymers 15:1345–1375
NM, Satyanarayana S (2012) Synthesis, characterization and 48. Satyanarayana S, Dabrowiak JC, Chaires JB (1992) Neither delta-
DNA-binding characteristics of Ru (II) molecular light switch nor lambda-tris (phenanthroline) ruthenium (II) binds to DNA by
complexes. J Iran Chem Soc 9:671–680 classical intercalation. BioChemistry 31:9319–9324
30. Srishailam A, Kumar YP, Reddy PV, Nambigari N, Vuruputuri U, 49. Barton JK, Danishefsky A, Goldberg J (1984) Tris (phenanthro-
Singh SS, Satyanarayana S (2014) Cellular uptake, cytotoxicity, line) ruthenium (II): stereoselectivity in binding to DNA. J Am
apoptosis, DNA-binding, photocleavage and molecular docking Chem Soc 106:2172–2176
studies of ruthenium (II) polypyridyl complexes. J Photochem 50. Satyanarayana S, Dabrowiak JC, Chaires JB (1993) Tris (phenan-
Photobiol B Biol 132: 111–23 throline) ruthenium (II) enantiomer interactions with DNA: mode
31. Nagababu P, Shilpa M, Latha JN, Bhatnagar I, Srinivas PN, and specificity of binding. BioChemistry 32:2573–2584
Kumar YP, Reddy KL, Satyanarayana S (2011) Synthesis, charac- 51. Chaires JB, Dattagupta N, Crothers DM (1982) Self-association
terization, DNA binding properties, fluorescence studies and toxic of daunomycin. BioChemistry 21:3927–3932
activity of cobalt (III) and ruthenium (II) polypyridyl complexes. 52. Cohen G, Eisenberg H (1969) Viscosity and sedimentation study
J Fluoresc 21:563–572 of sonicated DNA–proflavine complexes. Biopolymers 8:45–55

13

2130 J Fluoresc (2017) 27:2119–2130

53. Stevenson P, Sones KR, Gicheru MM, Mwangi EK (1995) Com- 55. Kundu S, Maity S, Bhadra R, Ghosh P (2011) Trans-
parison of isometamidium chloride and homidium bromide as dichlorobis(N-p-tolylpyridin-2-amine)palladium(II): synthesis,
prophylactic drugs for trypanosomiasis in cattle at Nguruman, structure, fluorescence features and DNA binding. Indian J Chem
Kenya. Acta Trop 59:77–84 50: 1443–9
54. Olmsted J III, Kearns DR (1977) Mechanism of ethidium bromide
fluorescence enhancement on binding to nucleic acids. BioChem-
istry 16:3647–3654

13

Vous aimerez peut-être aussi