You are on page 1of 17

G Model

CCR-111879; No. of Pages 17 ARTICLE IN PRESS


Coordination Chemistry Reviews xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


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

Review

Control and utilization of ruthenium and rhodium metal complex


excited states for photoactivated cancer therapy
Jessica D. Knoll, Claudia Turro ∗
Department of Chemistry and Biochemistry, The Ohio State University, Columbus, OH 43210, USA

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
1.1. Cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
1.1.1. Targeting DNA with metal complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
1.1.2. Cisplatin and its derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
1.2. Photodynamic therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2. Oxygen-dependent DNA modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3. Oxygen-independent DNA modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.1. Redox reactions with DNA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.2. Covalent DNA binding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.3. Photoactivated drug release . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

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

Article history: The use of visible light to produce highly selective and potent drugs through photodynamic therapy (PDT)
Received 23 February 2014 holds much potential in the treatment of cancer. PDT agents can be designed to follow an O2 -dependent
Received in revised form 15 May 2014 mechanism by producing highly reactive species such as 1 O2 and/or an O2 independent mechanism
Accepted 20 May 2014
through processes such as excited state electron transfer, covalent binding to DNA or photoinduced drug
Available online xxx
delivery. Ru(II)-polypyridyl and Rh2 (II,II) complexes represent an important class of compounds that
can be tailored to exhibit desired photophysical properties and photochemical reactivity by judicious
Keywords:
selection of the ligand set. Complexes with relatively long-lived excited states and planar, intercalating
Photodynamic therapy
Excited state ligands localize on the DNA strand and photocleave DNA through 1 O2 production or guanine oxidation by
Singlet oxygen the excited state of the chromophore. Photoinduced ligand substitution occurs through the population of
DNA photocleavage triplet metal centered (3 MC) excited states and facilitates covalent binding of the metal complex to DNA
DNA binding in a mode similar to cisplatin. Ligand photodissociation also provides a route to selective drug delivery.
The ability to construct metal complexes with desired light absorbing and excited state properties by
ligand variation enables the design of PDT agents that can potentially provide combination therapy from
a single metal complex.
© 2014 Elsevier B.V. All rights reserved.

Abbreviations: 5CNU, 5-cyanouracil; 9-EtG, 9-ethylguanine; 9-MeG, 9-methylguanine; bete, 3,6-dithiaoctane; biq, 2,2 -biquinoline; bpm, 2,2 -bipyrimidine;
bpte, 1,2-bis(phenylthio)ethane; bpy, 2,2 -bipyridine; CS, charge separated; dae, 1,2-dianilinoethane; DAP, 1,12-diazaperylene; dpp, 2,3-bis(2-pyridyl)pyrazine;
dppn, benzo[i]dipyrido[3,2-a:2 ,3 -h]quinoxaline; dppz, dipyrido[3,2-a:2 ,3 -c]phenazine; dpq, dipyrido[3,2-f:2 ,3 -h]quinoxaline; dpqp, pyrazino[2 ,3 :5,6]pyrazino[2,3-
f][1,10]phenanthroline; EC50 , half maximal effective concentration; en, ethylenediamine; EtBr, ethidium bromide; GMP, guanosine monophosphate; hat, 1,4,5,8,9,12-
hexaazatriphenylene; IL, intraligand; LC50 , median lethal concentration; MC, metal centered; MCR, multicellular resistance; Me2 bpy, 6,6 -dimethyl-2,2 -bipyridine;
Me2 dpq, 7,10-dimethylpyrazino[2,3-f][1,10]phenanthroline; MLCT, metal-to-ligand charge transfer; MMCT, metal-to-metal charge transfer; OC, open circular; PDT, pho-
todynamic therapy; phen, 1,10-phenanthroline; phpy− , deprotonated phenylpyridine; PI, phototoxicity index; PS, photosensitizer; PSS, polystyrene sulfonate; pydppz,
3-(pyrid-2 -yl)dipyrido[3,2-a:2 ,3 -c]phenazine; pydppn, 3-(pyrid-2 -yl)-4,5,9,16-tetraaza-dibenzo[a,c]naphthacene; pydppx, 3-(pyrid-2 -yl)-11,12-dimethyldipyrido[3,2-
a:2 ,3 -c]phenazine; pyr, pyrene; ROS, reactive oxygen species; SC, supercoiled; tap, 1,4,5,8-tetraazaphenanthrene; t Bu2 bpy, 4,4 -di-tert-butyl-2,2 -bipyridine; tpy,
2,2 :2 ,6 -terpyridine; ˚, quantum yield; , lifetime.
∗ Corresponding author. Tel.: +1 614 292 6708.
E-mail addresses: turro.1@osu.edu, turro@chemistry.ohio-state.edu (C. Turro).

http://dx.doi.org/10.1016/j.ccr.2014.05.018
0010-8545/© 2014 Elsevier B.V. All rights reserved.

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
2 J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx

1. Introduction ␲-stack. The complex can also bind in the major or minor groove
of the double helix based on size, shape, and their ability to hydro-
1.1. Cancer gen bond. Electrostatic attraction of metal complexes with cationic
charges aid in binding to DNA.
The pursuit of anti-cancer therapies that are highly effective
and exhibit low systemic toxicity is an important research area 1.1.2. Cisplatin and its derivatives
spanning many interdisciplinary fields. Cancer is defined as uncon- Cisplatin, cis-[Pt(NH3 )2 Cl2 ] [7–9], and its derivatives carboplatin
trolled proliferation of abnormal cells that accumulate to form and oxaliplatin [10] are the only metal-based drugs approved
tumors or lesions that interfere with the functioning of normal in the United States for cancer therapy to date, and they are
tissues and organs [1]. When the cancerous cells from a primary utilized in more than half of all cancer treatments, usually in
tumor migrate through the body to affect a non-adjacent organ or combination with radiation or other drugs [6,11–13]. Cisplatin
tissue, metastatic cancer develops. The three current cancer treat- undergoes thermal ligand exchange in aqueous solution to pro-
ments in clinical use are plagued by severe side effects: surgery or duce the mono-aqua complex [Pt(NH3 )2 (OH2 )Cl]+ , followed by a
resection is invasive and increases the risk of metastasis, radiation second ligand exchange to form the bis-aqua active species, cis-
therapy can cause radiation poisoning and increase the probability [Pt(NH3 )2 (OH2 )2 ]2+ . This species, by virtue of labile Pt OH2 bonds,
of developing secondary tumors, and chemotherapy drugs suffer binds primarily to the N7 on guanine bases and predominantly
from systemic toxicity due to a lack of selectivity [2,3]. While can- forms 1,2-GpG intrastrand crosslinks, resulting in a kinked DNA
cer death rates have declined by >1% per year between 1999 and double helix [14]. This Pt-G covalent binding results in the inhi-
2009 [2], a great need persists for better treatments that are min- bition of transcription and DNA replication which is the mode of
imally invasive and significantly less or non-toxic toward healthy action against tumors. Cisplatin is plagued by severe drawbacks,
cells. including its inability to distinguish between healthy and cancerous
cells and causing aggressive systemic toxicity [11]. Drug resistance
1.1.1. Targeting DNA with metal complexes through enhanced or overexpression of nucleotide excision repair
The transcription and replication of DNA are central to cell is another common limitation of cisplatin [15].
proliferation, so targeting the material involved in this process
is important in inhibiting the growth of cancerous tumors [1]. 1.2. Photodynamic therapy
DNA codes genetic information through the sequence of hydrogen-
bonded bases, adenine (A), cytosine (C), guanine (G), and thymine The use of visible light to activate a molecule and yield a potent
(T), with a double-helical secondary structure made up of two drug with spatial and temporal selectivity can be achieved through
anionic deoxyribose phosphodiester backbones as the scaffold [4]. photodynamic therapy (PDT) [16–18]. Ideal PDT candidates should
The DNA helical structure gives rise to minor and major grooves. be minimally toxic in the dark, absorb low energy visible light
Fig. 1(a) highlights the secondary structure of a B-DNA double helix within the therapeutic window of 600–900 nm to penetrate tissue,
and Fig. 1(b) shows the GC and AT base pairs. selectively accumulate in the tumor, and be amphiphilic to trans-
The complexity of the DNA double helical structure provides verse the cellular membrane [19]. A Jablonski diagram depicting
various types of sites for a metal complex to bind through different the photophysical processes involved in PDT is presented in Fig. 2.
modes [4–6]. The complex can form a covalent bond between the A ground state (1 GS) photosensitizer (PS) absorbs visible light to
metal and the Lewis basic phosphodiester backbone or the nitrogen populate the singlet excited state (1 ES) which typically deactivates
sites on the bases. Metal complexes with planar, aromatic ligands back to the 1 GS through fluorescence or through population of the
can bind via intercalation between adjacent base pairs, or through triplet excited state (3 ES) via intersystem crossing (ISC). Deacti-
insertion of the molecule at mismatched or abasic sites, both of vation of the 3 ES to the 1 GS by emission of a photon (radiative
which are driven by ␲–␲ interactions between the ligand and DNA decay, phosphorescence) or by the release of heat (nonradiative

Fig. 1. Schematic representation of (a) the components of the B-DNA double helix and (b) guanine-cytosine and adenine-thymine base pairs formed by hydrogen bonding.

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx 3

Fig. 2. Jablonski diagram representing the general scheme for PDT highlighting Type I, II, and III mechanisms.

decay, thermal) can then occur. If the 3 ES is relatively long-lived, manner similar to cisplatin (Section 3.2), and if the photodisso-
it can undergo three types of reactions of interest in PDT. Type I is ciated ligand is biologically active, light activated drug delivery
electron transfer to O2 or other oxygen-containing species in solu- can be achieved (Section 3.3). Comparing the DNA–metal complex
tion which eventually result in the generation of reactive oxygen interactions between studies performed in various laboratories
species (ROS). Type II reactivity is ascribed to energy transfer to is complicated by the factors that impact binding such as ionic
ground state 3 O2 to produce the very reactive 1 O2 , and Type III strength and pH of the buffer, concentration of metal complex
activity results from electron transfer from the excited state of the and DNA, and the type of DNA selected for the study. Discussion
sensitizer to cellular targets. PDT has been utilized in the treatment is limited to Ru(II)-polypyridyl and Rh2 (II,II) complexes that func-
of bladder, gastrointestinal, prostate, and gynecological lesions as tion under visible light irradiation. With respect to light activated
well as early stage, inoperable esophageal cancer, early or late stage drug delivery via ligand dissociation, discussion will focus on the
head and neck tumors, inoperable early central lung cancers, and delivery of organic drugs featuring nitrile functional groups. Other
dermatology [16,20–24]. Production of 1 O2 provides high selectiv- transition metals, such as Pt, Ir, and Mo, have also been studied for
ity and localization, as the typical lifetime of 1 O2 in a metabolically photoactivated chemotherapy, and they are reviewed elsewhere
healthy cell is ∼3 ␮s, resulting in an estimated intracellular diffu- [30]. A discussion of Ru and Rh metallo-intercalators and metallo-
sion distance of 2–4 × 10−6 cm2 /s. [25]. Photofrin® , composed of insertors for charge transport through DNA or mismatch targeting
hematoporphyrin and its oligomers, was approved by the FDA in is not included as they are well reviewed elsewhere [31,32]. The
1995 to treat esophageal, head, and neck tumors and functions by substantial body of work on photoactivated delivery of NO [33–35],
the Type II mechanism [26,27]. The drug absorbs strongly at 400 nm CO [36], and amino acids and neurotransmitters [37–41] from metal
due to the Soret band transition and in the 500–600 nm range aris- complexes is beyond the scope of the present review and is not
ing from the Q band transitions. This PDT agent is hindered by its discussed herein.
non-unity population of the 3 ES (quantum yield, ˚, of 0.83) which
limits the quantum yield of 1 O2 production (˚1 O ) to 0.65 [27]. 2. Oxygen-dependent DNA modification
2
The most aggressive and drug resistant tumors are hypoxic, which
limits the efficacy of PDT drugs that function by sensitizing 1 O2 Ru(II)-polypyridyl complexes are frequently utilized in the
[26,28]. To overcome the limitations of 1 O2 -generating PDT agents, development of metal-based PDT agents as they exhibit rich visi-
photochemotherapeutic agents that undergo photoinduced ligand ble light absorption and undergo population of relatively long-lived
dissociation to covalently bind DNA in a manner similar to cisplatin 3 MLCT (metal-to-ligand charge transfer) excited states with unit

and/or deliver a drug with spatiotemporal selectivity are recently quantum yields [42–44]. These 3 MLCT states provide convenient
under ardent investigation [29,30]. probes into excited state reactivity, are typically emissive, can be
Discussed herein are the applications of ruthenium and strongly oxidizing and/or reducing, and are typically long-lived and
rhodium complexes in PDT, through both O2 -dependent and O2 - efficiently generate 1 O2 [45]. The cationic nature of many transi-
independent mechanisms. To date, all approved PDT agents are tion metal complexes aids in the pre-association of the complex
organic molecules that sensitize 1 O2 production, and cisplatin and with the polyanionic DNA through electrostatic attraction, and the
its derivatives are the only approved metal-based cancer drugs ground and excited state properties are conveniently tunable as
for cancer therapy. Section 2 discusses Ru and Rh complexes they are dictated by the ligand set [43,46]. The prototypical visible
that produce 1 O2 which ultimately results in DNA photocleavage. light absorber, [Ru(bpy)3 ]2+ (bpy = 2,2 -bipyridine), Fig. 3, absorbs
The relatively long-lived metal-to-ligand charge transfer (3 MLCT) strongly between 400 and 500 nm due to Ru(d␲) → bpy(␲*) 1 MLCT
excited states of Ru(II)-polypyridyl complexes and long-lived transitions [42]. The 3 MLCT excited state of the complex is pop-
intraligand (3 IL) excited states of Ru(II) and Rh2 (II,II) complexes ulated rapidly with unit efficiency (˚ = 1), is relatively long-lived
are well suited for 1 O2 production, and ligand set variation allows (0.61 ␮s in H2 O and 1.00 ␮s in D2 O) [47,48], and undergoes
for more efficient association with DNA. Ru and Rh complexes that energy transfer to 3 O2 to form the highly reactive 1 O2 with
photodamage DNA in the absence of O2 are discussed in Section ˚ = 0.22 in air-saturated D2 O [49], thus photocleaving pBR322
3. Excited state complexes with strong oxidizing power undergo plasmid (irr > 450 nm) via the Type II PDT mechanism [46,50,51].
photoinduced redox reactions with guanine bases to cleave DNA [Ru(bpy)3 ]2+ , a weak electrostatic DNA binder, is limited as a PDT
(Section 3.1). Photoinduced ligand dissociation can be exploited agent by its inability to intercalate or covalently bind to DNA,
to produce a metal complex that covalently binds to DNA in a resulting in less targeted 1 O2 delivery. The 1,10-phenanthroline

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
4 J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 3. Structural representations of (a) the octahedral geometry of a tris-bidentate Ru(II) complex and (b) the bidentate NN ligands involved in the study of DNA photocleavage
through 1 O2 production.

(phen) analog, [Ru(phen)3 ]2+ (Fig. 3), has a longer excited state life- thymus DNA to a solution of each of the DAP complexes reveals
time (0.962 ␮s in H2 O and 1.1 ␮s in D2 O) [52] and subsequently hypochromic and bathochromic shifts consistent with intercalation
is a more efficient 1 O2 producer in D2 O (˚ = 0.24) [49]. While (binding constants, Kb , of 1.4 − 1.6 × 106 M−1 ). The poor solubil-
[Ru(phen)3 ]2+ was initially believed to interact with DNA via inter- ity of [Ru(bpy)(DAP)2 ]2+ and [Ru(DAP)3 ]2+ in H2 O and the high
calation of the phen ligand, the currently accepted binding mode is concentrations necessary for thermal denaturation and relative vis-
its semi-intercalation [53,54]. The incorporation of the tridentate cosity studies precluded their analysis by these methods. However,
2,2 :2 ,6 -terpyridine (tpy) ligand to provide the [Ru(tpy)2 ]2+ ana- [Ru(bpy)2 (DAP)]2+ increases the melting temperature (Tm ) of calf
log results in a drastically shortened excited state lifetime of 0.12 ns thymus by +6 ◦ C, a value similar to that of the known interca-
due to the distorted octahedral geometry promoting population of lator ethidium bromide (EtBr), with Tm = +5 ◦ C [56]. Moreover,
the nonemissive 3 MC state (Fig. 4). This excited state does not per- changes in the relative viscosity of herring sperm DNA that par-
sist long enough to effectively sensitize 1 O2 production or cause allels EtBr were measured in the presence of [Ru(bpy)2 (DAP)]2+ .
DNA photodamage. [Ru(tpy)2 ]2+ electrostatically binds to the DNA These observations suggest a primary intercalative binding mode
polyanion in a manner similar to [Ru(bpy)3 ]2+ [46]. The inability for the complex afforded by the presence of the DAP ligand. Photo-
to absorb lower energy visible light in the therapeutic window is a cleavage of 100 ␮M pUC18 supercoiled plasmid DNA occurs when
limitation for these homoleptic chromophores. 20 ␮M [Ru(bpy)2 (DAP)]2+ is irradiated with  > 395 nm for 30 min
A series of Ru(II) complexes incorporate 1,12-diazaperylene in the air, and no photocleavage occurs when the solution is deoxy-
(DAP), a bpy analog with an extended ␲-system and larger sur- genated. The involvement of 1 O2 production in the photoreactivity
face area to facilitate more avid DNA binding via intercalation with DNA is supported by the increase in photocleavage in D2 O
[55]. The series features one, two, or three DAP ligands coor- compared to H2 O.
dinated around the Ru(II) center resulting in [Ru(bpy)2 (DAP)]2+ , The “DNA light-switch” complex [Ru(bpy)2 (dppz)]2+
[Ru(bpy)(DAP)2 ]2+ , and [Ru(DAP)3 ]2+ (Fig. 3). The complexes are (dppz = dipyrido[3,2-a:2 ,3 -c]phenazine), Fig. 3, is an exten-
strong UV and visible light absorbers, with the Ru(d␲) → DAP(␲*) sively studied molecule that intercalates between DNA base pairs
1 MLCT transitions centered at 552 nm, 576 nm, and 588 nm in via the rigid, planar dppz ligand [57–59]. The light switch effect
CH3 CN for complexes with one, two, and three DAP ligands, respec- refers to the enhancement of the luminescence of the excited state
tively. The red shift that results from additional DAP ligands is species in the presence of DNA, since the complex in non-emissive
accompanied with an increase in molar absorptivity, owing to the in aqueous media in the absence of DNA, but its luminescence
strong absorption of transitions involving DAP in comparison to “turns on” upon intercalating. [Ru(bpy)2 (dppz)]2+ possesses two
those of bpy. The absorption of [Ru(DAP)3 ]2+ extends to 675 nm, low-lying excited states: a dark (non-emissive) 3 MLCT state involv-
an important characteristic in developing therapeutic agents for ing the phenazine moiety of dppz distal to the metal and a bright
PDT. The complexes are nonemissive at room temperature (RT) in (emissive) 3 MLCT state involving the proximal bpy portion of the
CH3 CN, and at 77 K in EtOH/MeOH glass weak emission is observed dppz ligand. In aqueous solution, the dark state is lower in energy
between 698 and 727 nm. The weak luminescence is attributed to than the bright state; the latter not thermally accessible at RT, such
the deactivation of the 3 MLCT state by a low-lying DAP dark 3 ␲␲* that the complex is non-emissive. In organic solvents or when
intraligand (IL) state. The interactions of these compounds with intercalated into DNA, the dark state energy is raised in energy
DNA were evaluated by electronic absorption titrations, thermal closer to the bright state energy, allowing thermal population and
denaturation, and relative viscosity changes. The addition of calf enhanced emission. The ability of this complex to participate in

Fig. 4. Structural representations of (a) the octahedral geometry of a bis-tridentate Ru(II) complex and (b) the tridentate NNN ligands involved in the study of DNA
photocleavage through 1 O2 production.

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx 5

intercalative binding to DNA is useful in PDT applications. The and 1.2-fold, respectively. The difference can be attributed to the
impact of substituted dppz-type ligands on the intercalation and notion that only one pydppz ligand on [Ru(pydppz)2 ]2+ interca-
light-switch effect has been observed for a variety of complexes lates between the DNA base pairs while the other pydppz remains
[60]. The ability of [Ru(bpy)2 (dppz)]2+ to photocleave DNA in the unbound and exposed to solvent. It is expected that the lowest
presence of O2 was compared to an analog, [Ru(bpy)2 (dpqp)]2+ energy 3 MLCT excited is localized on the non-intercalated lig-
(dpqp = pyrazino[2 ,3 :5,6]pyrazino[2,3-f][1,10]phenanthroline; and, such that its surroundings are not affected significantly by
Fig. 3), which exhibits unusually strong emission in H2 O at RT [58]. the intercalation of the other pydppz ligand, such that the com-
Both complexes possess similar absorption characteristics, with plex exhibits very little change in emission. When irradiated with
the Ru → L 1 MLCT transition centered at 445 nm (16,300 M−1 cm−1 ) irr > 395 nm for 10 min in the presence of O2 and pUC18 plasmid,
and 457 nm (12,300 M−1 cm−1 ) for L = dppz and dpqp, respectively. only [Ru(tpy)(pydppz)]2+ causes photocleavage. This photocleav-
The dpqp complex is strongly emissive in aqueous solution at age occurs only in the presence of O2 despite the short 5 ns lifetime
617 nm (˚em = 0.039,  = 582 ns). These values are quite similar and a low ˚1 O of 1.9 ± 0.3%. The shorter excited state lifetime of
2
to those of [Ru(bpy)3 ]2+ (em = 626 nm, ˚em = 0.042, 630 ns). [Ru(pydppz)2 ]2+ is consistent with its lower ˚1 O of 0.9 ± 0.4% and
2
Although the photophysical properties are not well understood, its lack of DNA photocleavage.
the emission of the dpqp complex is proposed to arise from a To extend the excited state lifetimes of the [Ru(tpy)(pydppz)]2+
3 MLCT excited state localized on the bpy portion of the dpqp
and [Ru(pydppz)2 ]2+ complexes while maintaining the inter-
ligand. The binding constant for [Ru(bpy)2 (dpqp)]2+ with DNA calative binding mode, a series of analogous homoleptic
was measured as 2.0 × 106 M−1 (s = 1.62), which is similar to those and heteroleptic compounds with extended ␲-systems
of [Ru(bpy)2 (dppz)]2+ , 106 –107 M−1 , indicating that the former were constructed using the ligands pydppx (3-(pyrid-2 -yl)-
likely undergoes intercalative binding in a manner similar to the 11,12-dimethyldipyrido[3,2-a:2 ,3 -c]phenazine) and pydppn
latter. The DNA photocleavage ability of these complexes was (3-(pyrid-2 -yl)-4,5,9,16-tetraaza-dibenzo[a,c]naphthacene)
probed with pUC18 plasmid DNA (irr > 455 nm, 15 min) in the (structures shown in Fig. 4) [62]. The lowest energy 1 MLCT absorp-
presence of O2 . A larger degree of photocleavage of the plasmid tion bands are similar to those of the analogous pydppz complexes,
occurs with [Ru(bpy)2 (dpqp)]2+ than with [Ru(bpy)2 (dppz)]2+ , however, the photophysical processes are greatly impacted when
consistent with the larger ˚1 O of [Ru(bpy)2 (dpqp)]2+ (0.76) pydppn is employed. In the complexes containing dppn ligands,
2
compared to [Ru(bpy)2 (dppz)]2+ (0.16). An even larger ˚1 O of the lowest energy excited state is 3 ␲␲* in nature rather than
2
3 MLCT, as is the case for the pydppz and pydppx complexes.
0.81 was reported for [Ru(bpy)3 ]2+ , but very little photocleavage
occurs under these conditions because it does not intercalate and This low-lying ligand-centered excited state is long lived with
it is only a weak electrostatic binder. In all cases, no photocleavage  = 20.1 and 24.3 ␮s for [Ru(tpy)(pydppn)]2+ and [Ru(pydppn)2 ]2+ ,
occurs in the absence of O2 . respectively. These values are in stark contrast to the short 3 MLCT
The intercalating dppz moiety was incorporated into the lifetimes of the aforementioned pydppz complex and those incor-
[Ru(tpy)2 ]2+ architecture to provide a means of DNA binding porating pydppx, with  = 3.7 for [Ru(tpy)(pydppx)]2+ and 1.1 ns
and extend the excited state lifetime to promote 1 O2 produc- for [Ru(pydppx)2 ]2+ , as the 3 ␲␲* state is higher energy than the
3 MLCT state. The trend is consistent with the extended ␲-system
tion [57]. The heteroleptic [Ru(tpy)(pydppz)]2+ and homolep-
tic [Ru(pydppz)2 ]2+ (pydppz = 3-(pyrid-2 -yl)dipyrido[3,2-a:2 ,3 - and addition of electron donating substituents to the dppz motif.
c]phenazine), for which the MLCT absorption maxima are observed The long-lived excited states of the pydppn complexes result in
between 475 and 481 nm, are shown in Fig. 4. The visible light large ˚1 O values of 0.92(2) and 1.07(7) for [Ru(tpy)(pydppn)]2+
2
absorption features are quite similar to those of the parent and [Ru(pydppn)2 ]2+ , respectively. The excited state lifetimes and
[Ru(tpy)2 ]2+ complex despite the presence of the dppz moiety in ˚1 O values for the pydppx complexes, 0.014(3) and 0.010(4) for
2
pydppz, attributed to the optical transition taking place from the [Ru(tpy)(pydppx)]2+ and [Ru(pydppx)2 ]2+ , respectively, are similar
metal to the proximal bpy or tpy portion of the pydppz ligand. to those of the pydppz analogs. The remarkable ˚1 O values for the
2
Both complexes are weakly emissive at RT; [Ru(tpy)(pydppz)]2+ pydppn complexes are greater than that of hematoporphyrin and
exhibits an emission maximum at 698 nm (˚em = 0.00021) and represent the most efficient 1 O2 production to date. The binding
[Ru(pydppz)2 ]2+ at 678 nm (˚em = 0.00061) in CH3 CN. The emis- constants of the complexes with calf thymus DNA were deter-
sion of each pydppz complex is red shifted and more efficient than mined to be 4.6 × 106 M−1 for [Ru(tpy)(pydppn)]2+ , 6.9 × 105 M−1
that of [Ru(tpy)2 ]2+ (em = 629 nm, ˚em < 5 × 10−6 ), as the pyd- for [Ru(tpy)(pydppx)]2+ , and 3.5 × 105 M−1 for [Ru(pydppn)2 ]2+ . No
dpz ligand is involved in the emissive 3 MLCT state. The excited binding constant was measured for [Ru(pydppx)2 ]2+ due to aggre-
state lifetimes determined by transient absorption spectroscopy gation of the complex. The lower Kb value for [Ru(tpy)(pydppx)]2+
are 5 ns and 2.4 ns for [Ru(tpy)(pydppz)]2+ and [Ru(pydppz)2 ]2+ , may be due to steric hindrance of intercalation by methyl sub-
respectively. The short excited state lifetimes are due to thermal stituents. The pydppn complexes efficiently photocleave pUC18
population of the low-lying 3 MC state, typical of [Ru(tpy)2 ]2+ - plasmid DNA by an O2 mediated pathway when irradiated with
type complexes ( = 126 ps for [Ru(tpy)2 ]2+ ). The intercalation  > 395 nm for 10 min. Complete conversion to nicked, open
of the pydppz ligand is supported by the typical hypochromic circular plasmid occurs with [Ru(tpy)(pydppn)]2+ , consistent with
and bathochromic shifts in the electronic absorption spectra of single-strand breaks that result from the production of ROS and
the complexes upon addition of calf thymus DNA, a trend not guanine oxidation derived from 1 O2 . The photocleavage is not as
observed for [Ru(tpy)2 ]2+ . The binding constants, 2.0 × 106 M−1 efficient for [Ru(pydppn)2 ]2+ due to its lower Kb value, and the
and 8.1 × 106 M−1 for [Ru(tpy)(pydppz)]2+ and [Ru(pydppz)2 ]2+ , short excited state lifetimes and inefficient 1 O2 generation of the
respectively, are similar to those of [Ru(phen)2 (dppz)]2+ [61]. pydppx complexes result in insignificant DNA photocleavage.
Intercalation was confirmed for [Ru(tpy)(pydppz)]2+ by relative The efficient O2 mediated DNA photocleavage by
viscosity measurements with herring sperm DNA, but the low sol- [Ru(tpy)(pydppn)]2+ and [Ru(pydppn)2 ]2+ motivated the study of
ubility of [Ru(pydppz)2 ]2+ in buffer did not allow the same analysis, the impact of these complexes on photocrosslinking of Proliferat-
although it is expected to exhibit similar behavior as the heterolep- ing Cell Nuclear Antigen (PCNA) and p53 protein in mammalian
tic analog. Additionally, the change in emission intensity upon cells and cell lysates [63]. The protein p53 maintains genetic
addition of calf thymus DNA is quite different between the het- stability and is involved in signaling pathways necessary for cell
eroleptic and homoleptic complexes, with enhancements of 8-fold survival, aging, and cancer formation, prognosis, and treatment.

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
6 J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 5. A series of Rh2 (II,II)-dppn complexes that photocleave DNA by a predominantly O2 -dependent mechanism.

More than half of all cancers exhibit mutations of the homote- with polyanionic polystyrene sulfonate (PSS) in optical titrations.
trameric p53. PCNA, a circular homotrimer that encircles one The complexes with NN = bpy and phen intercalate between
ds-DNA, is vital in mammalian replication and DNA repair. These base pairs as NN is not long enough to block dppn intercalation
nuclear targets are important in the treatment of tumors. Pho- (Kb = 1.2 × 106 and 1.0 × 105 M−1 , respectively). The dpq complex
tocrosslinking studies were performed with African green monkey intercalates and aggregates on the DNA surface, while the dppz
kidney fibroblasts (CV-1) and human fibroblasts (GM639). While and dppn complexes interact with DNA solely through surface
[Ru(tpy)(pydppn)]2+ and [Ru(pydppn)2 ]2+ differ in their ability aggregation, enhanced by the polyanionic DNA or PSS structure.
to photocrosslink p53 and PCNA in cells (white light, 3.15 J/cm2 , When irradiated in the presence of pUC18 plasmid DNA in an air
7 min) with the heteroleptic complex more efficiently inducing saturated solution ( > 375 nm, 20 min), the bpy, phen, dpq, and
crosslinks, the two complexes exhibit similar efficiencies in cell dppz complexes convert 64–68% of the supercoiled (SC) plasmid
lysates. This result suggests a difference in ability of the complexes to the open circular (OC) form, while dppn only converts 43%.
to enter the cells or to reach the nucleus. [Ru(tpy)(pydppn)]2+ The first four complexes undergo a dominant O2 -dependent and
also induces protein–DNA crosslinks and inhibits DNA replication minor O2 -independent photocleavage mechanism, as removal of
when irradiated, suggesting that DNA damage signaling and cell O2 decreases activity but does not completely eliminate it. The
cycle checkpoint pathways remain functional after damage to removal of O2 has no impact on the activity of the dppn complex.
nuclear proteins. As the dppz and dppn complexes produce 1 O2 with similar effi-
A series of cis-[Rh2 (-O2 CCH3 )2 (dppn)(NN)]2+ complexes, ciencies, the difference in activity is unexpected; however, the
where NN = bpy, phen, dpq = dipyrido[3,2-f:2 ,3 -h]quinoxaline, complex with one dppz and one dppn ligand may be able to allow
dppz, and dppn = benzo[i]dipyrido[3,2-a:2 ,3 -h]quinoxaline partial intercalation of dppn for more efficient ROS delivery, while
(Fig. 5), possess long-lived dppn 3 ␲␲* states and to photocleave that with two dppn ligands cannot intercalate.
DNA with a dominant O2 -dependent mechanism; the complexes The cis-[Rh2 (-O2 CCH3 )2 (dppn)(NN)]2+ (NN = bpy, phen, dpq,
in the series exhibit varying degrees of toxicity toward HeLa dppz, and dppn) complexes exhibit dark toxicity toward HeLa and
and COLO-316 cancer cell lines [64]. The electronic absorption COLO-316 cancer cell lines, with the bpy, phen, dpq, and dppz com-
spectrum of each complex features ligand-centered ␲␲* transi- plexes transversing the cellular membrane to induce apoptosis,
tions between 316 and 329 nm associated with both the NN and while the dppn complex is unable to enter the cell and induces
dppn ligands. The MLCT transitions are expected to occur with necrosis [65]. The cytotoxicity studies of this series toward Hs-
lower molar extinction coefficients between 400 and 430 nm, 27 fibroblasts shows that the dppz and dppn complexes are less
hidden beneath the broad 1 ␲␲ * dppn transitions, and weak cytotoxic in the dark than the bpy, phen, and dpq complexes, with
Rh2 (␲*) → Rh2 (␴*) MC bands are observed in the 600–621 nm LC50 values of 200 ± 20, 178 ± 17, 51 ± 5, 355 ± 18, and 384 ± 24 ␮M
range. No 3 MLCT emission is detected in H2 O for the cis-[Rh2 (- (LC50 = median lethal concentration) for NN = bpy, phen, dpq, dppz,
O2 CCH3 )2 (dppn)(NN)]2+ complexes, consistent with the transient and dppn, respectively [64]. The enhanced hydrophobicity of the
absorption spectroscopy results which reveal that the long-lived, NN = dppz and dppn complexes (log P = 0.62 and 1.02, respectively)
lowest energy excited state is dppn 3 ␲␲* in nature, with lifetimes suggests that they may not cross through the cellular membrane
of 2.7, 2.4, 2.4, 3.5, and 4.1 ␮s for NN = bpy, phen, dpq, dppz, and as well as those with NN = bpy, phen, and dpq (log P = 0.3, 0.30,
dppn, respectively. However, the respective 1 O2 quantum yields, and 0.41, respectively). Upon irradiation, the complexes exhibit
˚1 O , measured to be 0.7, 0.9, 0.8, 0.4, and 0.4, do not follow the a greater potency toward the cells, with LC50 values of 30 ± 5,
2
trend expected based on the lifetimes. The lower 1 O2 production 22 ± 4, 9 ± 3, 17 ± 3, and 16 ± 4 ␮M for NN = bpy, phen, dpq, dppz,
efficiency for the dppz and dppn complexes, despite their longer- and dppn, respectively. Whereas complexes with NN = bpy, phen,
lived excited states, is attributed to the greater hydrophobicity of and dpp may enter into the cell and damage intracellular materials
these two complexes relative to others in the series, which may via ROS generation, the more hydrophobic complexes may remain
cause aggregation in H2 O. The complexes were determined to bound within the cellular membrane and be photoreactive toward
either intercalate or aggregate in the presence of DNA (or both), phospholipids to induce cell death. The increase in toxicity upon
the aggregation was established through the replacement of DNA irradiation, between 5.7-fold and 8.4-fold, for NN = bpy, phen, and

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx 7

Fig. 6. Structural representations of (a) the octahedral geometry of the Ru(II) complexes and (b) the ligands utilized in complexes that undergo photoinduced redox reactions
with DNA.

dpq are similar to the increase for hematoporphyrin, while L = dppz that intersystem crossing from the initially generated singlet
and dppn Exhibit 21-f old and 24-fold enhancements. Along with state populates both the 3 MLCT and 3 ␲␲* states. As expected
their lower dark toxicity than hematoporphyrin, these complexes based its long-lived LC excited state, [Ru(bpy)2 (dppn)]2+ more
represent potential improvements over current PDT drugs. efficiently produces 1 O2 (˚ = 0.88 ± 0.05) than the dppz analog
(˚ = 0.16 ± 0.02). The DNA binding constant of [Ru(bpy)2 (dppn)]2+ ,
3. Oxygen-independent DNA modification Kb , is on the order of 106 M−1 , similar to those of other known
DNA intercalators, and the increased relative viscosity of herring
3.1. Redox reactions with DNA sperm DNA upon addition of complex confirms intercalation as the
primary binding mode of the complex. Complete and rapid pho-
A method for O2 -independent DNA photocleavage involves tocleavage of pUC18 plasmid DNA (100 ␮M bases) occurs when
redox reactions with DNA, primarily guanine oxidation by an irradiated with  > 455 nm for 30 s, while no cleavage is observed for
excited state species [66]. The [Ru(t Bu2 bpy)2 (Mebpy-V1 -V2 )]6+ [Ru(bpy)3 ]2+ or [Ru(bpy)2 (dppz)]2+ under these conditions. Signif-
complex (t Bu2 bpy = 4,4 -di-tert-butyl-2,2 -bipyridine) with two icant photocleavage occurs within 3 min of irradiation of the dppn
viologens tethered to the Ru(II)-polypyridyl chromophore (Fig. 6) complex with  > 550 nm. In D2 O only a slight enhancement of DNA
was designed for charge separation (CS) between the spatially photocleavage is observed compared to the system in H2 O, suggest-
separated HOMO (localized on Ru) and LUMO (localized on V2 ). ing very little impact of 1 O2 on the reactivity with the duplex. A
The electronic absorption spectrum of this complex features a small amount of cleavage occurs when the sample is deoxygenated
1 MLCT absorption maximum at 449 nm (11,200 M−1 cm−1 ), typical and in the presence of NaN3 (a 1 O2 or • OH scavenger) or superox-
of Ru(II)-polypyridyl complexes. Very weak emission is observed ide dismutase (SOD), indicating that an additional, O2 -independent
at 641 nm (˚em = 0.0050(5), N2 purged H2 O), which is quenched by mechanism plays a role in photocleavage. This mechanism involves
O2 . A long-lived CS state ( = 1.7 ␮s) is produced upon excitation guanine oxidation by the excited state complex. The excited state
with  = 532 nm as observed by transient absorption spectroscopy, reduction potential (Ered *) is approximately +1.64 V vs NHE, mak-
which features peaks typical of reduced viologen at 395 and ing it a strong oxidizing agent. Guanine oxidation in H2 O requires
605 nm. The [Ru(t Bu2 bpy)2 (Mebpy-V1 -V2 )]6+ complex (3 ␮M) pho- +1.29 V vs NHE at neutral pH, providing a thermodynamically favor-
tocleaves pUC18 plasmid DNA (100 ␮M) in the presence and able process (G = −0.35 V). Stern-Volmer quenching of the excited
absence of O2 when irradiated with  > 395 for 20 min. This state [Ru(bpy)2 (dppn)]2+ upon addition of guanosine monophos-
O2 -independent cleavage occurs via guanine oxidation by the long- phate (GMP) yields a quenching constant (kq ) of 2.1 × 108 M−1 s−1 ,
lived photogenerated Ru(III) center in the excited state species. This which represents a relatively efficient process. No quenching is
effect is similar to Ru(II) tris-polypyridyl complexes in the presence observed for [Ru(bpy)3 ]2+ by GMP, since guanine oxidation from its
of an electron acceptor [67,68]. excited state is not thermodynamically favorable. While the Os(II)
[Ru(bpy)2 (dppn)]2+ was investigated as an analog to the analog, [Os(bpy)2 (dppn)]2+ , absorbs in the therapeutic window and
O2 -dependent DNA photocleavage agent [Ru(bpy)2 (dppz)]2+ pre- photocleaves DNA when irradiated with red light ( > 645 nm for
viously discussed. However, [Ru(bpy)2 (dppn)]2+ cleaves plasmid 1.5 h) in the presence of O2 , it does not photocleave DNA under
DNA in the absence of O2 [69]. The electronic absorption spec- hypoxic conditions as the excited state species is too low in energy
trum of [Ru(bpy)2 (dppn)]2+ features dppn 1 ␲␲* transitions at to oxidize guanine [70]. It is believed that the [Ru(bpy)2 (dppn)]2+
387 nm (9900 M−1 cm−1 ) and 411 nm (13,400 M−1 cm−1 ) and a complex undergoes interesting dual activity with DNA through a
1 MLCT absorption centered at 444 (13,500 M−1 cm−1 ). The spec- highly reactive MLCT state that oxidizes guanine and a long-lived
3 ␲␲* state that produces ROS.
tral features of this complex in the visible light range are quite
similar to those of [Ru(bpy)3 ]2+ and [Ru(bpy)2 (dppz)]2+ . A weak Polypyridyl ligands structurally related to phen, such
emission is observed at 617 nm (˚em = 0.003) from a 3 MLCT excited as tap (1,4,5,8-tetraazaphenanthrene) and hat (1,4,5,8,9,12-
state that is relatively long-lived ( = 803 ns). Transient absorption hexaazatriphenylene), were incorporated into Ru(II) complexes
spectroscopy reveals population of a lowest energy, non-emissive to facilitate intercalation and photoinduced electron transfer to
excited state that is dppn-localized 3 ␲␲* in nature that is long oxidize guanine bases [43,51,71]. The homoleptic and heteroleptic
lived ( = 33 ␮s). Ultrafast transient absorption studies indicate DNA cleaving agents in this series are [Ru(tap)3 ]2+ , [Ru(hat)3 ]2+ ,

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
8 J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 7. Rh2 (II,II)-polypyridyl complexes that photocleave DNA in the absence of O2 .

[Ru(tap)2 (NN)]2+ , and [Ru(hat)2 (NN)]2+ (NN = bpy or phen; Fig. 6). feature associated with pyrene at 382 nm is observed in each
The complexes absorb efficiently with 1 MLCT transitions between complex. [Ru(bpy)2 (3-pyr-phen)]2+ and [Ru(bpy)2 (5-pyr-phen)]2+
400 and 500 nm (ε = 12,000–19,000 M−1 cm−1 ) [71–75]. Thermal exhibit similar absorption profiles with maxima near 412 nm and a
denaturation experiments and hypochromicity upon DNA addition lower energy shoulder at 450 nm, both corresponding to 1 MLCT
are consistent with the planar ligands partially intercalating transitions. The 1 MLCT absorption of [Ru(bpy)2 (4-pyr-phen)]2+
between the base pairs. Quenching of the 3 MLCT emission was is red shifted to 482 nm as substitution at the para position to
observed upon addition of calf thymus DNA for all members of the coordinated N atom of phen stabilizes the phen acceptor
this series. This behavior is the opposite of that observed for DNA orbitals and increases the communication between the pyrene
light-switch complexes, suggesting that the excited state reactiv- moiety and the Ru center. The complexes are weakly emissive
ity is quite different. Intermolecular electron transfer from GMP (em = 609–625 nm, ˚em < 0.005) in air saturated CH3 CN, with a
(E◦ (G•+ /G) = +0.92 V vs SCE) to the excited state of each complex lowest-energy excited 3 ␲␲* state localized on the pyrene ligand.
is thermodynamically favorable (Ered * = +1.46 V and +1.32 V for [Ru(bpy)2 (3-pyr-phen)]2+ and [Ru(bpy)2 (5-pyr-phen)]2+ have sim-
[Ru(hat)3 ]2+ and [Ru(tap)3 ]2+ , respectively; +1.23 V and +1.12 V ilar excited state lifetimes of 240 and 140 ␮s in deoxygenated
for [Ru(hat)2 (phen)]2+ and [Ru(hat)2 (bpy)]2+ , respectively; and CH3 CN, while [Ru(bpy)2 (4-pyr-phen)]2+ has a shorter lifetime of
+1.06 V for both [Ru(tap)2 (phen)]2+ and [Ru(tap)2 (bpy)]2+ ), such 22 ␮s, but the three complexes produce 1 O2 with almost identical
that photocleavage of pBR322 plasmid (irr = 436 nm) via guanine efficiency (˚ = 0.65, 0.68, and 0.67 for the 3-, 4-, and 5-substituted
oxidation is a likely mechanism. In contrast, [Ru(tap)(bpy)2 ]2+ complexes, respectively). The 3 MLCT excited state of each com-
and [Ru(tap)(phen)2 ]2+ , with Ered * = +0.83 V and +0.87 V vs SCE, plex is also populated upon excitation but the lifetimes at RT could
respectively, do not have sufficient excited state oxidizing power not be resolved with the ∼20 ns instrument response. Binding
to participate in guanine oxidation, and consequently very little titrations of [Ru(bpy-d8 )2 (3-pyr-phen)]2+ and [Ru(bpy-d8 )2 (5-pyr-
DNA photocleavage was observed. phen)]2+ with herring sperm DNA resulted in Kb values of 3.0 × 107
A series of dyad complexes with Ru(II)-polypyridyl chro- and 8.5 × 106 M−1 , respectively, consistent with those of known
mophores linked to a planar, rigid pyrene ligand through intercalators [78,79]. Thermal denaturation experiments with calf
an ethynyl linker were studied for their ability to photo- thymus DNA (pH 7.4) in the presence of the two complexes reveals
cleave DNA. The photocytotoxicity of the complexes toward a marked increase in the melting temperatures, with Tm values of
human leukemia cell lines (HL-60) and metastatic melanoma +20 and +10 ◦ C, respectively, relative to DNA alone. The difference
cell line Malme-3M was also investigated as a function of the in melting temperature is sensitive to the placement of the pyrene
placement of the pyrenylethynylene substituent on the phen substituent around the phen ligand. Both complexes substantially
ligand [76,77]. The [Ru(bpy)2 (3-pyr-phen)]2+ , [Ru(bpy)2 (4-pyr- increased the DNA viscosity, another confirmation of intercalation.
phen)]2+ , and [Ru(bpy)2 (5-pyr-phen)]2+ complexes (Fig. 6) feature The two bpy-d8 complexes photocleave pUC19 plasmid DNA when
a [Ru(bpy)2 (phen)]2+ parent complex with the pyrenylethyny- irradiated with 420 nm light for 30 min with half maximal effective
lene substituent at the 3-, 4-, and 5-position on the phen ligand, concentration (EC50 ) values in the nanomolar range, and com-
respectively. The compounds exhibit strong IL and MLCT absorp- plete conversion to OC plasmid occurs with complex concentration
tion in the ultraviolet and visible spectral regions, and the as low as 2 ␮M. A plasmid with greater GC content, pDesR3, is

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx 9

linearized through double-strand photocleavage with concentra- of the DNA polyanion enhances intermolecular ␲-stacking of the
tions as low as 3 ␮M for the 3-substituted complex, while no complex. The three complexes photocleave pUC18 plasmid DNA
linearization occurs with the 5-substituted analog. The photocleav- upon irradiation with  > 395 nm for 15 min, but the photocleav-
age occurs via an O2 -independent mechanism and it is proposed age by Rh2 -mono-dppz is greater due to its ability to intercalate.
to take place through guanine oxidation by the excited state com- In contrast, the homoleptic cis-[Rh2 (-O2 CCH3 )2 (bpy)2 ]2+ and cis-
plex. This conclusion is supported by more avid photocleavage with [Rh2 (-O2 CCH3 )2 (phen)2 ]2+ complexes do not photocleave DNA
increasing guanine content in the plasmid. [84,85]. The DNA photocleavage of Rh2 -mono-dppz slightly dimin-
The cytotoxicity and photocytotoxicity of these complexes were ishes in the absence of O2 , but the O2 -independent mechanism is
first assayed with a human leukemia cell line, HL-60, which showed substantial. It is suggested that a lowest lying 3 MLCT excited state of
that they have high phototoxicity index (PI) values, defined as the Rh2 -mono-dppz, Rh2 -bis-dppz, and Rh2 -bpy-dppz, in which elec-
ratio of the dark EC50 to the irradiated EC50 . A species such as tron density moves from the dirhodium core to the phenazine
cisplatin, which is not photochemically active, has a PI equal to moiety of the dppz ligand, results a longer-lived charged-separated
1. The PI values for the 3-, 4-, and 5-substituted complexes are excited state relative to those of the bpy and phen analogs, allow-
104, 382, and 1747, respectively (photolyzed with 100 J/cm2 white ing sufficient time for the partially oxidized dirhodium core to
light for 1 h), with irradiated EC50 values of 1.8, 0.22, and 0.15 ␮M, cleave DNA. These complexes were evaluated for their cytotox-
respectively. The 5-pyr-phen complex exhibits a very low toxicity icity and photocytotoxicity toward Hs-27 cells. When incubated
in the dark, with EC50 = 262 ␮M, which when combined with its with the cells in the dark, Rh2 -mono-dppz, Rh2 -bis-dppz, and Rh2 -
low nanomolar irradiated EC50 value, results in an unprecedented bpy-dppz have LC50 values of 27 ± 2 and 135 ± 8, and 208 ± 10 ␮M,
PI that is an order of magnitude greater than the largest reported respectively. It is believed that the intercalation of the Rh2 -mono-
to date (vide infra) [80]. A lower light dosage of 5 J/cm2 still results dppz complex may contribute to its increased dark toxicity. When
in 0.5–1.5 ␮M EC50 values, and the cell death occurs via apopto- irradiated ( > 400–700 nm, 30 min), the LC50 values decreased to
sis. It is important to note that the irradiated complexes are up 21 ± 3, 39 ± 1, and 44 ± 2 ␮M for Rh2 -mono-dppz, Rh2 -bis-dppz,
to 170 times more potent than cisplatin and are 10–85 times less and Rh2 -bpy-dppz, respectively. While the mono-dppz complex
toxic in the dark than cisplatin, making these compounds potential exhibits almost no difference between dark and irradiated cyto-
PDT agents with low systemic toxicity in the dark and very high toxicity, the Rh2 -bis-dppz and Rh2 -bpy-dppz complexes represent
cytotoxicity when irradiated. The activity of the complexes against good PDT candidates due to low dark toxicity and increased pho-
the metastatic melanoma cell line Malme-3M was also investigated totoxicity by factors that range from 3.4 to 4.7. In this motif,
in the dark and irradiated conditions. Malme-3M is an aggressive intercalation or strong DNA binding may not be desirable prop-
melanoma with a poor prognosis, no curative therapy, a relatively erties in developing successful PDT agents. The dirhodium species
high recurrence rate of resected melanoma, and a median survival have promising potential when compared to the toxicities of these
rate of only 9 months at stage IV. Because of its very low O2 content, currently employed cancer therapeutics.
it is untreatable by compounds that produce ROS [81]. Irradiation Recently, an unbridged Rh2 (II,II) complex with phen ligands
of the cells in the presence of each of the three complexes with that chelate to each Rh center, [Rh2 (phen)2 (CH3 CN)6 ]4+ (Fig. 7),
white light (7 J/cm2 , 15 min) resulted in cell death with nanomolar was investigated for its ability to photocleave DNA when exposed
EC50 values. The 5-pyr-phen complex exhibited an irradiated EC50 to visible light [86]. Upon low energy visible light photoly-
value of 200 nM and those for the 3- and 4-substituted complexes sis ( > 590 nm) in H2 O, the complex releases two equatorial
were measured to be in the 600–700 nM range. The dark toxicity CH3 CN ligands and undergoes homolytic Rh-Rh bond cleav-
of the complexes toward the Malme-3M cell line is slightly greater age, resulting in the formation of two monometallic radical
than observed for the HL-60 cells, with EC50 values of 62, 44, and Rh(II) fragments, which then generate a mononuclear aqua com-
57 ␮M for the 3-, 4-, and 5- pyr-phen complexes, respectively, but plex (˚aq = 1.38). The greater than unity quantum yield can be
the PI values are still quite large. The trends observed in the photo- indicative of an initial photoreaction followed by a dark reac-
cytotoxicity cannot be explained by the photophysical properties tion, or may be related to the formation of two mononuclear
of the complexes. Factors such as subcellular localization, biological rhodium complexes from the bimetallic reactant. The two major
targets, cellular uptake, metabolism, and mechanism of action are products observed by ESI-MS are [Rh(phen)(CH3 CN)(OH)]+ and
all important in cytotoxicity, and differences in the size, shape, and [Rh(phen)(CH3 CN)(OH2 )3 (BF4 )]+ . Formation of the reactive radical
hydrophobicity of the complexes are expected to impact toxicity. Rh(II) fragments under irradiation with  > 590 nm photocleaves
It is evident, however, that the combination of the low-lying 3 ␲␲* DNA in an O2 -independent manner, while this activity does not
state and the intercalating ability and lipophilicity of the pyrene lig- occur in the dark. This complex represents a potentially useful
and is important for activity. This series of potent photocytotoxic photo-therapeutic candidate owing to it reactivity with low energy
agents with low dark cytotoxicities suggests that strong intercala- light in the PDT window.
tion does not always induce dark cytotoxicity and this method of
targeting DNA should not be overlooked. 3.2. Covalent DNA binding
In addition to the ruthenium-based systems, the acetate-
bridged Rh2 (II,II) compounds depicted in Fig. 7 incorporate Dirhodium (II,II) and oxidized dirhodium(III,II) paddlewheel
bidentate ligands with extended ␲-systems that impart inter- complexes with four acetate bridges, [Rh2 (-O2 CCH3 )4 ] and
esting photophysical properties and intercalative abilities on the [Rh2 (-O2 CCH3 )4 ]+ (Fig. 8) exhibit anti-tumor activity against
complexes [82,83]. The two complexes with one or two dppz L1210 and Ehrlich ascites tumors in mice several decades ago
ligands, cis-[Rh2 (-O2 CCH3 )2 (dppz)(␩1 -O2 CCH3 )(CH3 OH)]+ (Rh2 - [87–90]. It is believed that these complexes act by binding to DNA
mono-dppz) and cis-[Rh2 (-O2 CCH3 )2 (dppz)2 ]2+ (Rh2 -bis-dppz) and inhibiting replication. Coordinate bond formation between
behave differently with DNA. Rh2 -mono-dppz binds to DNA by the axial positions and ligands containing O, N, S, or P donor
intercalation of the dppz ligand and it also exhibits a small atoms (typically solvent molecules) occurs readily. The Rh2 (II,II)
amount of surface aggregation. In contrast, Rh2 -bis-dppz and the complexes have long-lived excited states (typically 3.5–5.0 ␮s),
heteroleptic analog cis-[Rh2 (-O2 CCH3 )2 (bpy)(dppz)]2+ (Rh2 -bpy- however no photocleavage of pUC18 plasmid occurs with visible
dppz) do not intercalate, as the binding constant for the latter light irradiation ( > 395 nm). Upon addition of electron acceptors,
of 2.8 × 103 M−1 determined by equilibrium dialysis is inconsis- such as 3-cyano-1-methylpyridinium or 1,8-anthraquinone disul-
tent with intercalation (Kb ∼ 105 –106 M−1 ). However, the presence fonate, oxidation of the complex occurs to produce the Rh2 (III,II)

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
10 J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 8. Rh2 (II,II) paddlewheel complexes that undergo photoinduced covalent binding with DNA.

species which cleaves pUC18 plasmid with low energy excitation which is significantly less potent than when the complex is irra-
( > 610 nm), an important characteristic in development of PDT diated LC50 = 12 ± 2 ␮M (400–700 nm for 30 min). The PI of 34 is a
agents [85]. In the presence of O2 , the activity is greatly decreased large improvement over hematoporphyrin, with a PI measured to
suggesting that the excited state is readily deactivated by O2 . Addi- be 5.5 under similar experimental conditions. The dark LC50 value
tion of a non-labile ligand such as pyridine or PPh3 which bind to of hematoporphyrin (21 ± 1 ␮M) makes it ∼20 times more toxic
the axial sites in place of the labile OH2 prohibits photocleavage, as cis-[Rh2 (-O2 CCH3 )2 (CH3 CN)4 (OH2 )2 ]2+ in the dark. No photocy-
the ligand in the axial position must be labile in order to promote totoxicity is observed with cis-[Rh2 (-O2 CCH3 )2 (phen)2 (OH2 )2 ]2+
covalent interactions between one of the Rh atoms and DNA. Cova- as the phen ligands are not photolabile. Additionally, cis-[Rh2 (-
lent binding of adenine and guanine bases to this complex has been O2 CCH3 )2 (CH3 CN)4 (OH2 )2 ]2+ inhibits transcription by the T7-RNA
observed by mass spectrometry. A major drawback of this system polymerase enzyme with a Cinh 50 value of 2.6 ␮M, showing that the
is that thermal axial ligand exchange occurs in the dark. complex is more effective than cisplatin under similar conditions
The first reported metal-metal bonded complex to cova- (4.1 ␮M) [92].
lently bind to ds-DNA under visible light irradiation is of Ru(II)-polypyridyl complexes were designed that are inert in the
the form cis-[Rh2 (-O2 CCH3 )2 (CH3 CN)4 (OH2 )2 ]2+ , where the dark and covalently bind to DNA in a manner similar to cisplatin
water molecules are labile and occupy the axial positions only under visible light irradiation, whereby combining cisplatin’s
[91]. The cis-[Rh2 (-O2 CCH3 )2 (CH3 CN)6 ]2+ complex (Fig. 8), activity toward biological targets with the spatial and tempo-
which exhibits relatively weak absorption centered at 363 ral selectivity of PDT. Photoaquation of cis-[Ru(bpy)2 (NH3 )2 ]2+
and 525 nm (420 and 218 M−1 cm−1 , respectively), undergoes upon visible light irradiation produces cis-[Ru(bpy)2 (OH2 )2 ]2+ in
rapid thermal axial ligand exchange in H2 O in the dark to acidic solution and cis-[Ru(bpy)2 (OH2 )(OH)]+ in aqueous solu-
form cis-[Rh2 (-O2 CCH3 )2 (CH3 CN)4 (OH2 )2 ]2+ . The lowest energy tion at neutral pH (˚ = 0.024 with  = 350 nm; ˚ = 0.018 with
metal-centered Rh2 (␲*) → Rh2 (␴*) absorption shifts from 525  = 400 nm). The photoinduced ligand exchange is believed to pro-
to 555 nm (160 M−1 cm−1 ), as expected for a transition that ceed via thermal population of the low-lying 3 MC excited state
involves the Rh2 (␴*) orbital, such that its energy is sensitive that promotes monodentate ligand dissociation [93]. This pho-
to the nature of the axial ligands. Photolysis of the complex toactive cisplatin analog binds covalently to oligonucleotides via
with  > 455 nm for 5 h results in the exchange of two equatorial substitution of the thermally labile Ru OH2 bonds. The electronic
CH3 CN ligands with solvent H2 O molecules to yield cis-[Rh2 (- absorption spectrum of cis-[Ru(bpy)2 (NH3 )2 ]2+ in H2 O exhibits
O2 CCH3 )2 (CH3 CN)2 (OH2 )4 ]2+ . This photoaquation causes a shift of Ru(d␲) → bpy(␲*) 1 MLCT transitions at 345 nm (7340 M−1 cm−1 )
the transitions formerly centered at 373 and 555 nm to 450 and and 490 nm (8210 M−1 cm−1 ), and the complex is weakly emissive
573 nm, as predicted by time-dependent density functional the- in EtOH/MeOH glass at 157 K (em = 741 nm, ˚ = 0.002,  = 52 ns)
ory (TD-DFT) calculations. The two remaining CH3 CN ligands can and non-emissive in H2 O at RT. Sequential dissociation of the two
occupy different equatorial positions, resulting in three possible NH3 ligands is observed as evidenced by the biphasic changes to
isomers of the complex, and it is unclear which is formed. The the electronic absorption spectrum as a function of irradiation
quantum yield of photoaquation (˚aq ) is dependent on irradia- time as the complex is photolyzed in H2 O under an argon atmo-
tion wavelength with values of 0.37 and 0.09 for irr = 355 and sphere. The photoaquation product undergoes thermal reactions
509 nm, respectively. The power dependence of product formation with 9-methylguanine (9-MeG) and 9-EtG to produce the corre-
is consistent with a one photon process, such that the absorp- sponding [Ru(bpy)2 (9-MeG)]2+ and [Ru(bpy)2 (9-EtG)]2+ adducts.
tion of a single photon causes one CH3 CN substitution and the The bis-aqua product also covalently binds to a single-stranded 15-
second CH3 CN substitution must occur by a dark reaction. Upon mer, and enhanced binding is observed with a strand containing a
irradiation with  > 455 nm, the complex covalently binds to bpy, GpG step, one of the known preferential binding sites for cisplatin.
9-ethylguanine (9-EtG), and linearized pUC18 plasmid. This cova- Intrastrand covalent binding was demonstrated by the decrease in
lent binding does not occur in the dark. The complex’s ability to kill the melting temperature upon irradiation, Tm = −5 ◦ C, of the dou-
Hs-27 human skin cells upon irradiation and in the dark was inves- ble stranded 15-mer in the presence of the complex, as interstrand
tigated. In the dark, the complex has an LC50 value of 410 ± 9 ␮M, crosslinks destabilize the duplex and decrease the DNA melting

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx 11

Fig. 9. Ru(II)-polypyridyl complexes that undergo photoinduced ligand release from a RuII (bpy)2 fragment that covalently binds to DNA.

temperature. For comparison, cisplatin causes an 8 ◦ C decrease with low affinity [80]. Photoaquation of both complexes to produce
in the melting temperature of a double stranded 20-mer. More- cis-[Ru(bpy)2 (OH2 )2 ]2+ occurs when irradiated with  > 450 nm.
over, the decreased electrophoretic mobility of linearized pUC18 This ligand dissociation, causing a red shift in the MLCT maximum
plasmid in agarose gels upon irradiation in the presence of the com- from 450 to 480 nm, proceeds rapidly for the Me2 bpy complex
plex ( > 345 nm, 15 min) is also consistent with covalent binding. with t1/2 = 2 min, which is enhanced by a factor of 30 compared
The photoactive cis-[Ru(bpy)2 (NH3 )2 ]2+ represents an improve- to the analogous ligand exchange in the Me2 dpq complex. It is
ment over the thermally activated analogs cis-[Ru(bpy)2 (OH2 )2 ]2+ suggested that the rigidity of the planar dpq ligand facilitates
and cis-[Ru(phen)2 Cl2 ] species, since activation with light provides its re-coordination after one Ru-N bond is cleaved, as a stepwise
spatial and temporal control of reactivity [94,95]. dissociative bond breaking process leads to ligand exchange.
A related series of photoactive [Ru(bpy)2 (LL)]2+ complexes [Ru(bpy)2 (Me2 bpy)]2+ covalently binds to pUC19 upon irradiation,
with a variety of bidentate ligands, LL, featuring either N- or and the Me2 dpq complex both covalently binds and photocleaves
S-coordination to the metal were investigated to gain further the plasmid (white light, 200 W, 1 h). The reactivity with DNA
understanding of the dominant factors associated with light driven is unaffected by addition of glutathione (GSH), although GSH
ligand exchange from Ru(II) polypyridyl complexes [96]. The coordinates to and inhibits the activity of Pt drugs.
complexes [Ru(bpy)2 (LL)]2+ , where LL = bete = 3,6-dithiaoctane, The cytotoxicity of [Ru(bpy)2 (Me2 bpy)]2+ and
bpte = 1,2-bis(phenylthio)ethane, en = ethylenediamine, and [Ru(bpy)2 (Me2 dpq)]2+ under dark and irradiated conditions
dae = 1,2-dianilinoethane (Fig. 9), exhibit visible light absorption was investigated with HL-60 leukemia cells and A549 lung cancer
dictated by the nature of the ligand LL. In the S-coordinated cells. The complexes are more potent upon irradiation than cis-
complexes with LL = bete and bpte, the lowest energy 1 MLCT platin and are virtually nontoxic in the dark, exhibiting PI values
absorption maxima are observed at 422 and 404 nm, respectively, of 100–200. They also cause complete cell death when irradiated,
while these transitions in the N-containing complexes, LL = en a common difficulty for PDT agents. The dark IC50 values for
and dae, are red-shifted to 485 and 469 nm, respectively. Analysis [Ru(bpy)2 (Me2 bpy)]2+ are >300 ␮M (HL60) and 150(7) ␮M (A549)
of the photoaquation efficiency (irr = 400 nm) for the N-bound and those for [Ru(bpy)2 (Me2 dpq)]2+ are 108(2) ␮M (HL60) and
complexes was precluded by an overlap in spectra of the reac- 250(5) ␮M (A549). Upon irradiation with  > 450 nm (410 W,
tant and product, but [Ru(bpy)2 (bete)]2+ and [Ru(bpy)2 (bpte)]2+ 3 min), the IC50 values decreased significantly for both complexes.
undergo photoaquation with ˚aq = 0.024(2) and 0.022(3), respec- Values of 1.6(2) ␮M (HL60) and 1.1(3) ␮M (A549) were reported for
tively. The complexes were photolyzed in CH2 Cl2 with excess [Ru(bpy)2 (Me2 bpy)]2+ and 2.6 ± 1.0 ␮M (HL60) and 1.2 ± 0.1 ␮M
n-tetrabutylammonium chloride (TBACl) to compare the ligand (A549) for [Ru(bpy)2 (Me2 dpq)]2+ . A549 3D tumor spheroids
substitution efficiencies for the formation of [Ru(bpy)2 Cl2 ]. The provide a more accurate model to approximate the complexity
N-containing ligands undergo photosubstitution almost an order of in vivo tumors [97]. The complexes exhibit even lower dark
of magnitude lower than the corresponding S-coordinated ligands, toxicity with A549 spheroids, with LC50 values >300 ␮M for both
with ˚ = 0.019(1), 0.016(3), 0.002(1), and 0.003(1) for LL = bete, complexes. The irradiated IC50 values, 21.3 ± 2.3 and 64.6 ± 4.7 ␮M
bpte, en, and dae, respectively. The Ru-L bond elongation in the for [Ru(bpy)2 (Me2 bpy)]2+ and [Ru(bpy)2 (Me2 dpq)]2+ , respectively,
3 MLCT state was calculated to be greater with bete and bpte than are an order of magnitude larger than the monolayer models yet
with en and dae, a factor that likely plays a role in the enhanced still effective. The multicellular resistance ratio (MCR) is a ratio
ligand dissociation in the former. A thermal dark reaction between of spheroid to monolayer IC50 values. It should also be noted
GMP and the photogenerated [Ru(bpy)2 (OH2 )2 ]2+ complex from that the complexes have MCR values of 19.4 and 54; the value
the bete complex produces a (bpy)2 RuII -GMP adduct. The com- for [Ru(bpy)2 (Me2 bpy)]2+ is similar to the MCR value of 12.4 for
plexes with phenyl-containing ligands (bpte and dae) covalently cisplatin, which is lower than most chemotherapy drugs.
bind to linearized pUC18 upon irradiation ( > 395 nm, 4 min) more Following the principles of the methyl-substituted bpy and
efficiently than those containing bete and en, while the S-bound dpq ligands to distort the octahedral geometry in order to pro-
species are better DNA binders than the N-bound complexes, mote ligand dissociation, complexes featuring the sterically
consistent with the trends in photoaquation yields. demanding ligand 2,2 -biquinoline (biq), [Ru(phen)2 (biq)]2+ and
The photoaquation of Ru(II)-polypyridyl complexes is enhanced [Ru(phen)(biq)2 ]2+ (Fig. 10), were investigated [98]. Employing
through the distortion of the octahedral geometry with the the biq ligand in octahedral Ru(II)-polypyridyl complexes also
introduction of sterically bulky bidentate ligands in the coor- serves to red-shift the visible light absorption into the therapeutic
dination sphere, which promotes more efficient 3 MC state window because biq has lower-lying acceptor orbitals. The lowest
population. The two complexes [Ru(bpy)2 (Me2 bpy)]2+ and energy absorption of [Ru(phen)2 (biq)]2+ is centered at 525 nm
[Ru(bpy)2 (Me2 dpq)]2+ (Me2 bpy = 6,6 -dimethyl-2,2 -bipyridine; with appreciable absorption extending to 700 nm. The absorption
Me2 dpq = 7,10-dimethylpyrazino[2,3-f][1,10]phenanthroline; is even further red shifted for [Ru(phen)(biq)2 ]2+ , with a lowest
Fig. 9) were compared to [Ru(bpy)3 ]2+ , which lacks the sterically energy absorption centered at 550 nm and a tail extending to
bulky ligand and binds to DNA through electrostatic interactions 800 nm. One biq ligand is released upon photolysis in H2 O in each

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
12 J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 10. Ru(II)-polypyridyl complexes employing biq that covalently bind to DNA via photoinduced ligand release.

complex, resulting in the formation of the corresponding bis-aqua (X = halide). The halide undergoes ligand exchange by a thermal
species. The irradiation time required to generate the product process, which results in cytotoxicity toward several cancer cell
is an order of magnitude longer for [Ru(phen)2 (biq)]2+ than for lines, although it lacks spatial and temporal selectivity. The [(p-
[Ru(phen)(biq)2 ]2+ as a result of enhanced octahedral distortion cym)Ru(NN)(L)]2+ series represent the first reported examples of
from an additional bulky biq ligand in the latter. The complexes Ru(II) arene complexes that selectively release a monodentate lig-
do not induce photocleavage of pUC19 plasmid, but the bis-aqua and and covalently bind to DNA upon irradiation with visible light.
photolysis products covalently bind to DNA, reducing the mobility The electronic absorption spectrum of [(p-cym)Ru(bpm)(py)]2+ in
on agarose gel in a manner similar to cisplatin. The complexes H2 O features maxima at 254 and 383 nm, and these transitions red-
are proposed to bind to the duplex via intrastrand crosslinking. shift upon photolysis, consistent with dissociation of py to form
No binding occurs in the dark, and the mobility is similar to the [(p-cym)Ru(bpm)(OH2 )]2+ in aqueous media. Similar shifts were
synthesized cis-[Ru(phen)2 (OH2 )2 ]2+ . An irradiation wavelength observed for [(p-cym)Ru(phen)(py)]2+ and [(p-cym)Ru(bpm)(4,4 -
dependence shows the greatest activity with blue light, yet irra- bpy)]2+ , for which the low energy absorption tail is composed of a
diation with red wavelengths still induces significant covalent mixture of transitions to 1 MC and 1 MLCT excited states. The lowest
binding to DNA. These complexes are also potent toward HL-60 energy excited state has significant 3 MC character, consistent with
human leukemia cells, with IC50 values of 1–2 ␮M under blue light the lack of emission in solution and the observed photodissocia-
for 3 min, similar to the toxicity of cisplatin with IC50 = 3.1 ␮M. tion of the py ligand. Irradiation of [(p-cym)Ru(bpm)(py)]2+ in the
The potency decreases under red light but longer irradiation times presence of 9-EtG in aqueous solution forms [(p-cym)Ru(bpm)(9-
up to 6 min provides IC50 values of 2.3–7.6 ␮M. The PI values EtG)]2+ . This trend is consistent for the other two complexes
under blue irradiation are 43.8 and 19.7 for [Ru(phen)2 (biq)]2+ and of this motif, however, 9-ethyladenine (9-EtA) does not bind
[Ru(phen)(biq)2 ]2+ , respectively. to the photolyzed complex, highlighting the selectivity toward
A common disadvantage to using [Ru(bpy)3 ]2+ -type complexes guanine. While these complexes do not exhibit major differences
for PDT applications is the need for high energy visible light in their photochemistry, the reactivity with DNA differs upon
excitation as these species typically do not absorb appreciably structural changes. [(p-cym)Ru(bpm)(4,4 -bpy)]2+ undergoes 20%
in the therapeutic window. To shift the visible light absorp- DNA binding in the dark after 48 h, potentially via intercalation of
tion of photoactive Ru(II)-polypyridyl cisplatin analogs into the pendant 4,4 -bpy ligand, whereas <5% of [(p-cym)Ru(bpm)(py)]2+
PDT window, the deprotonated phenylpyridine (phpy− ) ligand and [(p-cym)Ru(phen)(py)]2+ bind to DNA over the same period
was incorporated to destabilize the Ru(d␲) orbitals in cis- of time. The binding is more rapid when the complex is irradiated
[Ru(phpy)(phen)(CH3 CN)2 ]+ (Fig. 11) [99]. Additionally, CH3 CN prior to the addition of DNA as compared to the irradiation of
was employed as the photolabile ligand, as the photoaqua- solutions that are composed of the complex and DNA, but the
tion of nitriles is generally much more efficient than that overall degree of binding is not impacted. The DNA binding of
observed for amines or pyridine [93,100]. The lowest energy elec- the photoproduct is slow compared to that of cisplatin; this is
tronic absorption peaks of the complex are centered at 461 nm an attractive feature for PDT drugs because low dark reactivity
(ε = 11,559 M−1 cm−1 ) and 486 nm (ε = 10,842 M−1 cm−1 ), corre- with biomolecules is desired to prevent unwanted side reactions.
sponding to Ru(d␲) → phen(␲*) 1 MLCT. More importantly, the Upon irradiation in the presence of SC DNA, the complexes cause
complex has an appreciable absorption tail extending beyond 6–7◦ unwinding, similar to that observed with monofunctional
600 nm into the therapeutic window. Upon irradiation with cisplatin adducts [105]. Additionally, the complexes inhibit DNA
 > 420 nm for 20 min, cis-[Ru(phpy)(phen)(CH3 CN)2 ]+ covalently and RNA transcription upon irradiation, and mapping experiments
binds to linearized plasmid DNA, evidenced by slower migration in determined that the stop sites are always guanine bases. Despite
the gel mobility assay. The complex inhibits tumor growth in mice the lack of ligand dissociation in the dark, the complexes exhibit
as well as cytotoxicity with light exposure, however detailed photo- dark cytotoxicity toward the A2780 human ovarian cancer cell
physical or photochemical studies were not reported [101,102]. The line, however, the mechanism remains unknown.
complex exhibits dark and red-light induced cytotoxicity toward Two new complexes inspired by the photoreactive piano-stool
OVCAR-5 cells (human advanced ovarian epithelial cancer cell line), Ru(II)-arene complexes were designed to incorporate receptor
with a dark LC50 value of 1.0 ␮M. This represents a fairly high dark binding peptides for tumor-targeted drug delivery. The peptides
toxicity and is attributed to the enhanced thermal ligand exchange are incorporated through the py ligand in [(p-cym)Ru(bpm)(py)]2+ .
in the presence of GSH, a molecule that is found at relatively high The two selected peptides are the dicarba analog of octreotide, a
intracellular concentrations. The irradiated complex ( = 690 nm, potent endocrine hormone somatostatin agonist containing eight
100 s), however, is quite potent with an LC50 value of 70 nM, repre- peptides, and the Arg-Gly-Asp (RGD) sequence, a moiety that
senting a 14-fold increase in toxicity upon red light exposure well selectively targets tumor endothelial cells over healthy cells. The
into the PDT window. receptors for each of the peptides are overexpressed in various
A series of photoactive piano-stool type com- tumors and have been employed in cytotoxic drug delivery [106].
plex, [(p-cym)Ru(NN)(L)]2+ (p-cym = para-cymene; The complexes do not undergo ligand substitution in the dark,
NN = bpm = 2,2 -bipyrimidine or phen; L = py = pyridine or 4,4 - but irradiation with blue light (420 nm) at 37 ◦ C results in the
bpy = 4,4 -bipyridine; Fig. 12) [103,104] was designed to improve exchange of py-substituted peptide with a solvent H2 O molecule,
upon the previously reported [(p-cym)Ru(bpm)(X)]n+ complex evident by the observation of the free substituted peptide and the

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx 13

Fig. 11. Ru(II)-polypyridyl complexes utilizing phpy− that covalently bind to DNA via photoinduced ligand release.

aqua product [(p-cym)Ru(bpm)(OH2 )]2+ by 1 H NMR and electronic [{(bpy)2 Ru(dpp)}2 RhCl2 ]5+ , [{(bpy)2 Os(dpp)}2 RhCl2 ]5+ ,
absorption spectroscopy for each complex. Both complexes form [{(tpy)RuCl(dpp)}2 RhCl2 ]3+ , and [{(bpy)2 Ru(bpm)}2 RhCl2 ]5+
adducts with 9-EtG when irradiated in H2 O in a manner simi- (dpp = 2,3-bis(2-pyridyl)pyrazine; Fig. 13), were studied to under-
lar to the parent [(p-cym)Ru(bpm)(py)]2+ . When irradiated for 8 h stand the impact of component variation on the biological activity
followed by the addition of the single-stranded oligonucleotide 5 - [107–111]. The HOMO in each complex is localized on the Ru(II)
dCATGGCT-3 and overnight incubation, the octreotide complex or Os(II) center. The LUMO is localized on the Rh center for the
covalently binds to either of the guanine bases. When irradiated dpp-bridged complexes as the Rh(␴*) orbitals are lower energy
in the presence of the oligonucleotide for 9 h, the two monofunc- than the dpp(␲*) orbitals; however, in the bpm-bridged complex
tional adducts were observed in addition to a bifunctional adduct it is centered on one of the bpm ligands, since the bpm(␲*) orbitals
formed by dissociation of p-cym to form a more stable GG chelate. are lower lying than the Rh(␴*) orbitals. The electronic absorption
When a similar experiment was performed with an oligonucleotide spectra feature Ru/Os(d␲) → dpp/bpm(␲*) 1 MLCT transitions
containing two non-adjacent G bases, 5 -dAGCCATG, the two centered at 530 nm for the dpp complexes and 594 nm for the bpm
monofunctional adducts and a cyclic adduct by GG chelation fol- complex, as bpm(␲*) orbitals are stabilized compared to dpp. The
lowing arene dissociation were formed. A peptide-oligonucleotide Os(II) complex also absorbs appreciably in the therapeutic window

hybrid, Phac-His-Gly-Met-linker-p5 -dCATGGCT, was employed to due to the direct spin-forbidden transition, 1 GS → 3 MLCT, afforded
investigate the competition between DNA and protein binding. This by the spin-orbit coupling induced by the third-row transition
peptide was chosen because histidine and methionine are the two metal. Upon irradiation with  > 475 nm, the dpp-bridged com-
residues most likely to form complexes with Ru. Irradiation for 9 h plexes covalently bind to and photocleave pUC18 and pBluescript
results in no Met or His binding, but monofunctional G adducts DNA in the absence of O2 . The excited state complex rapidly
were observed, indicative of the selectivity for DNA. The p-cym populates a Ru/Os(d␲) → dpp(␲*) 3 MLCT excited state followed
ligand only dissociated to enable GG chelation after extensive irra- by relaxation to the GS via radiative or nonradiative decay or
diation. population of a low-lying Ru/Os(d␲) → Rh(␴*) 3 MMCT (metal-to-
A supramolecular architecture coupling one or two Ru(II)- metal charge transfer state) via intramolecular electron transfer.
polypyridyl chromophores to a RhIII Cl2 DNA binding site exhibit In the 3 MMCT state the electron density localizes on the Rh center,
photophysical properties that strongly influence their ability causing labilization of the Rh(II) Cl bond and providing an open
to covalent bind and cleave DNA in the presence of light with- site for coordination of DNA bases. The bpm-bridged complex does
out O2 . A series of Ru(II),Rh(III),Ru(II) trimetallic complexes, not exhibit covalent DNA binding as the Ru(d␲) → bpm(␲*) 3 MLCT

Fig. 12. Ru(II)-arene complexes that covalently bind DNA in the absence of O2 .

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
14 J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 13. Ru(II),Rh(III) supramolecular complexes that undergo photoinduced covalent binding to DNA in the absence of O2 .

state is lower in energy than the 3 MMCT state, thus hindering efficient under irradiation in the therapeutic window, and cleave
electron transfer to the Rh center. The [{(bpy)2 Ru(dpp)}2 RhCl2 ]5+ DNA and inhibit replication makes these systems promising
and [{(bpy)2 Os(dpp)}2 RhCl2 ]5+ complexes, when irradiated with PDT agents.
 > 460 nm for 4 min, inhibit growth of African green monkey kid-
ney epithelial cells [108]. Interestingly, the Os complex inhibits cell 3.3. Photoactivated drug release
replication while the Ru complex kills cells in the presence of light.
Bimetallic Ru(II),Rh(III) analogs to the previously discussed The thermally accessible 3 MC excited states of Ru(II)-
Ru(II),Rh(III),Ru(II) trimetallic complexes were designed in an effort polypyridyl complexes to induce ligand dissociation is employed
to provide a more sterically accessible Rh center, better cell per- for light-activated drug delivery. Cancer therapy drugs and cysteine
meability, a lower overall charge, fewer stereoisomers, and a more protease inhibitors featuring nitrile substituents that form coordi-
easily tunable Rh(III) bioactive site [112–114]. The three complexes, nate bonds with a Ru(II) chromophore and are photolabile upon
[(bpy)2 Os(dpp)RhCl2 (phen)]3+ , [(bpy)2 Ru(dpp)RhCl2 (phen)]3+ , visible light excitation are discussed herein.
and [(bpy)2 Ru(bpm)RhCl2 (phen)]3+ exhibit similar photophys- The chemotherapy drug 5-fluorouracil (5FU) has been utilized
ical properties to the trimetallic systems. The Ru/Os(d␲) → for over 20 years and continues to be widely used in the treat-
dpp/bpm(␲*) 1 MLCT absorption occurs at 521, 515, and 592 nm, ment of malignancies, such as colorectal and breast cancers [115].
respectively. [(bpy)2 Os(dpp)RhCl2 (phen)]3+ absorbs broadly in The related molecule 5-cyanouracil (5CNU) inhibits pyrimidine
the NIR with a maximum at 750 nm due to direct population of catabolism in vivo [116], and the nitrile functional group enables it
the 3 MLCT state. Similar to the trimetallic complexes, the LUMO to coordinate to a Ru(II) photoactive unit. Two Ru(II) complexes, cis-
is localized on Rh in the dpp-bridged complexes and on bpm in [Ru(bpy)2 (5CNU)2 ]2+ and [Ru(tpy)(5CNU)3 ]2+ (Fig. 12), are able to
the bpm-bridged complex. Slower halide dissociation upon Rh photochemically deliver both the 5CNU drug and the Ru(II)–OH2
reduction is observed in electrochemical analysis of these systems compounds that covalently bind to DNA with spatial and tem-
compared to the trimetallic complexes. All three complexes bind poral control, making them potential dual activity chemotherapy
to and photocleave pUC18 when irradiated with 455 nm light, and agents [117,118]. The irradiation of cis-[Ru(bpy)2 (5CNU)2 ]2+ with
the activity with DNA is less efficient for the bpm complex. This is  > 395 nm forms cis-[Ru(bpy)2 (OH2 )2 ]2+ (˚aq = 0.16(4), 400 nm,
attributed to its shorter excited state lifetime (<10 ns) compared for the first 5CNU exchange), evidenced by the red shift in
to that of [(bpy)2 Ru(dpp)RhCl2 (phen)]3+ (30 ns). Further analysis the electronic absorption maximum from 410 nm to 450 nm at
of [(bpy)2 Ru(dpp)RhCl2 (phen)]3+ demonstrates that when irra- short irradiation times associated with the formation of the cis-
diated in NaH2 PO4 buffer with 455 nm light for 30 min, chloride [Ru(bpy)(5CNU)(OH2 )]2+ intermediate [117]. This step is followed
dissociation occurs as observed by a subtle blue shift in the 1 MLCT by a shift from 450 nm to 490 nm with longer irradiation times,
absorption. Mass spectrometry reveals that OH2 , OH− , and HPO4 − corresponding to formation of the bis-aqua product. The latter
bind to the open coordination site on Rh [114]. The Os complex is covalently binds to linearized pUC18 plasmid DNA with  > 395 nm
remarkably able to covalently bind to and fully photocleave DNA in irradiation for 15 min, and no covalent binding occurs when the
the absence of O2 with  > 645 nm for 4 h [112]. The complex also complex is incubated in the dark under similar experimental con-
inhibits DNA replication in PCR (polymerase chain reaction) assays ditions; moreover the mono-aqua complex does not bind to DNA.
using a 670 base pair fragment of pUC18. At a base pair-to-metal Cis-[Ru(bpy)2 (5CNU)2 ]2+ can deliver two equivalents of 5CNU and
complex ratio as low as 50:1, DNA replication is disrupted with one equivalent of a RuII (bpy)2 -binding agent to provide biologi-
red light irradiation for 1 h in the absence of O2 [113]. The ability cal activity that operates via two different mechanism, akin to the
of these mixed-metal systems to undergo 3 MMCT-mediated mixtures of drugs typically used for cancer treatment, commonly
halide dissociation to covalent bind to DNA, with some systems referred to as “drug cocktails.”

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx 15

to form the mono-aqua intermediate with ˚ = 0.080(4). Further


photolysis causes the substitution of the second NC-peptide to pro-
duce the bis-aqua [Ru(bpy)2 (OH2 )2 ]2+ product with ˚ = 0.011(1).
An investigation into the in vitro activity of the complex with the
cysteine protease papain shows that under irradiation ( > 395 nm,
10 min) the concentration required to inhibit 50% of the enzyme
activity, IC50 , has a value of 295 nM, which suggests that approxi-
mately two RCN ligands are released, since it is approximately half
the IC50 value for the free RCN ligand, 638 nM. Cysteine protease
inhibition is enhanced 32-fold when irradiated, based on a dark
IC50 of 9.5 ␮M. This low level of dark toxicity may stem from a
small amount of RCN dissociating under the reaction conditions
or the complex may undergo nonbonding interactions with the
enzyme that cause minor inhibition. The IC50 for the analogous
[Ru(bpy)2 (CH3 CN)2 ]2+ complex is >500 ␮M in the dark or irradi-
ated, providing strong evidence that the released RCN ligand, and
not the metal complex, is in fact the enzyme inhibitor. The complex
Fig. 14. Ru(II) complexes that deliver nitrile-containing drugs upon irradiation. is less potent for cathepsins B, K, and L, yet significant enhance-
ment of inhibition is observed upon irradiation compared to the
dark incubation.
The tpy analog [Ru(tpy)(5CNU)3 ]2+ was utilized to extend visible
light absorption toward the therapeutic window relative to that of
the cis-[Ru(bpy)2 (5CNU)2 ]2+ complex [118]. Additionally, RuII (tpy) 4. Conclusions
complexes bind 9-EtG [119] and [Ru(tpy)Cl3 ] is more cytotoxic
than [Ru(bpy)2 Cl2 ] [120]. The electronic absorption spectrum of PDT provides avenues for improving upon currently employed
[Ru(tpy)(5CNU)3 ]2+ in H2 O exhibits an MLCT absorption centered at chemotherapy drugs by producing potent drugs with spatial and
420 nm and a broad, low energy tail that extends to almost 600 nm, temporal selectivity determined by visible light irradiation. Ru(II)
compared to just beyond 500 nm for cis-[Ru(bpy)2 (5CNU)2 ]2+ . Pho- and Rh2 (II,II) complexes are attractive candidates for success-
tolysis with  > 395 in H2 O results in a red shift of the absorption ful PDT agents and are under investigation due to the ability to
maximum from 420 to 450 nm, consistent with the exchange of tune the excited states toward more efficient low-energy vis-
one axial 5CNU ligand with a solvent H2 O molecule. The substitu- ible light absorption, relatively long lifetimes, and control the
tion of the second axial 5CNU, evidenced by a further red shift of the type of reactivity to be achieved. Complexes with RuII (NN)2 and
absorption band to 475 nm, occurs with significantly longer irradi- RuII (NNN) fragments coordinated to planar ligands localize on DNA
ation times. The overall ˚ from reactant to product was measured by intercalation between base pairs, and can produce 1 O2 with
to be 0.022(2) (irr = 400 nm), a process that results in covalent quantum efficiency. The 1 O2 producing Rh2 (II,II) motif, cis-[Rh2 (-
binding to linearized pUC19 plasmid ( > 395 nm). The complex is O2 CCH3 )2 (dppn)(NN)]2+ , exhibits a decreased dark toxicity and
not cytotoxic toward HeLa cells when incubated in the dark for an increased phototoxicity toward Hs-27 fibroblasts compared to
2 h, but irradiation for 1 h with  > 400 nm kills HeLa cells with hematoporphyrin. Related complexes with intercalating ligands
LC50 = 156 ± 18 ␮M. This value is similar to that measured for the damage DNA via excited state electron transfer to oxidize gua-
free 5CNU ligand, 151 ± 33, suggesting only one 5CNU is released nine, and from this series, [Ru(bpy)2 (5-pyr-phen)]2+ exhibits an
per complex under the photolysis conditions used for the cytotoxi- unprecedented PI of 1747 against human leukemia HL-60 cells and
city experiments. A control experiment with [Ru(tpy)(CH3 CN)3 ]2+ , is quite active against Malme-3M aggressive melanoma cells that
which releases axial CH3 CN ligands upon irradiation, does not are untreatable by ROS. In addition to 1 O2 production and excited
exhibit photocytotoxicity. Since the mono-aqua species is expected state redox reactions, photoactive Ru and Rh complexes act as cis-
to form under the same conditions of the cell studies, the lack of platin analogs that covalently bind to DNA only under irradiation,
cytotoxicity is consistent because [Ru(tpy)(CH3 CN)2 (OH2 )]2+ does such that they may overcome undesired effects on normal cells. The
not covalently bind to DNA. This result provides strong evidence population of 3 MC states facilitates ligand substitution to release
that the 5CNU ligand is solely responsible for the photocytotoxicity monodentate or bidentate ligands from a RuII (NN)2 fragment to
as only the mono-aqua Ru complex is formed. form the cisplatin analog, [Ru(NN)2 (OH2 )2 ]2+ , which covalently
The RuII (bpy)2 moiety was also employed for light-activated binds to DNA. Members of this class of complexes are active against
delivery of a cysteine protease inhibitor [121]. Cysteine proteases lung cancer and leukemia cell lines. Analogous complexes with
play a contributing role in tumor growth, migration, angiogenesis, a phpy− ligand replacing one diimine NN ligand effectively red
and metastasis, and they are overexpressed in many cancers, posi- shifts the absorption into the therapeutic window and is active
tioning them as important targets in cancer therapy [122]. Cysteine against an ovarian cancer cell line. A series of Ru(II) piano-stool
proteases are also critical for normal cell functioning, so controlled complexes undergoes photoaquation from release of a pyridine-
inhibition of these proteases is crucial for minimizing undesirable type ligand and subsequently bind to DNA covalently; appending
systemic toxicity. The coordination to a RuII (bpy)2 fragment by the photoreleased pyridine with oligonucleotides and peptides
a peptidomimetic nitrile-based inhibitor serves to inactivate the affords a method for targeting tumors. Bimetallic and trimetallic
inhibitor until it is released by visible light. This strategy provides complexes featuring one or two Ru(II) chromophores bridged to
high selectivity for cysteine protease inhibition, as the nitrile group a RhIII Cl2 bioactive site covalently bind to DNA due to Cl− disso-
is guarded from attack by a cysteine residue in the active site of ciation resulting from population of an MMCT excited state, and
the enzyme when it is coordinated to the RuII (bpy)2 fragment. Cis- DNA replication inhibition was observed under red light irradia-
[Ru(bpy)2 (NC-peptide)2 ]2+ , depicted in Fig. 14, absorbs strongly in tion. Ligand substitution via 3 MC state population is also exploited
the visible region with a maximum at 414 nm in H2 O. The complex to release nitrile-containing drugs such as 5CNU and cysteine pro-
is stable in the dark but upon irradiation ( > 395 nm) the MLCT tease inhibitors, while also producing Ru(II)-OH2 complexes that
band red shifts to 444 nm due to substitution of one NC-peptide covalently bind to DNA. In developing new Ru(II) an Rh2 (II,II)

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
16 J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx

complexes to overcome the current challenges of PDT, complexes [38] M. Salierno, E. Marceca, D.S. Peterka, R. Yuste, R. Etchenique, J. Inorg. Biochem.
that require short irradiation times at long wavelengths are ideal. 104 (2010) 418–422.
[39] L. Zayat, C. Calero, P. Alborés, L. Baraldo, R. Etchenique, J. Am. Chem. Soc. 125
The need for low energy visible light absorption necessitates low- (2003) 882–883.
ering the activation barrier for thermal population of the 3 MC state [40] L. Zayat, M.G. Noval, J. Campi, C.I. Calero, D.J. Calvo, R. Etchenique, Chem-
from the low-lying 3 MLCT state for complexes that undergo ligand biochem 8 (2007) 2035–2038.
[41] L. Zayat, M. Salierno, R. Etchenique, Inorg. Chem. 45 (2006) 1728–1731.
dissociation for covalent binding or drug delivery. The ability to [42] A. Juris, V. Balzani, F. Barigelletti, S. Campagna, P. Belser, A. von Zelewsky,
alter the excited states and biological activity of Ru(II) and Rh2 (II,II) Coord. Chem. Rev. 84 (1988) 85–277.
complexes makes these systems versatile as they can function by [43] J.M. Kelly, D.J. McConnell, C. OhUigin, A.B. Tossi, A.K.-D. Mesmaeker, A. Mass-
chelein, J. Nasielski, J. Chem. Soc. Chem. Commun. (1987) 1821–1823.
more than one mechanism to enhance their effect on tumor cells
[44] J.N. Demas, D. Diemente, E.W. Harris, J. Am. Chem. Soc. 95 (1973) 6864–6865.
while minimizing the impact on healthy cells. Additionally, explo- [45] C.P. Anderson, D.J. Salmon, T.J. Meyer, R.C. Young, J. Am. Chem. Soc. 99 (1977)
ration of the in vitro and in vivo activity of these complexes will 1980–1982.
[46] J.M. Kelly, A.B. Tossi, D.J. McConnell, C. OhUigin, Nucleic Acids Res. 13 (1985)
be critical to their successes as PDT drugs. Collaborative, interdis-
6017–6034.
ciplinary efforts are essential in realizing the potential of PDT drug [47] Y. Kaizu, H. Ohta, K. Kobayashi, H. Kobayashi, K. Takuma, T. Matsuo, J. Pho-
design to generate more effective and non-invasive cancer therapy. tochem. 30 (1985) 93–103.
[48] M.L. Fetterolf, H.W. Offen, J. Phys. Chem. 90 (1986) 1828–1830.
[49] D. Garcia-Fresnadillo, Y. Georgiadou, G. Orellana, A.M. Braun, E. Oliveros, Helv.
Chim. Acta 79 (1996) 1222–1238.
Acknowledgments
[50] J.W. Dobrucki, J. Photochem. Photobiol. B 65 (2001) 136–144.
[51] A.B. Tossi, J.M. Kelly, Photochem. Photobiol. 49 (1989) 545–556.
The authors gratefully acknowledge support from the National [52] W.J. Dressick, B.L. Hauenstein, T.B. Gilbert, J.N. Demas, B.A. DeGraff, J. Phys.
Science Foundation (CHE-1213646) and the National Institutes of Chem. 88 (1984) 3337–3340.
[53] H. Niyazi, J.P. Hall, K. O’Sullivan, G. Winter, T. Sorensen, J.M. Kelly, C.J. Cardin,
Health (RO1 EB016072). Nat. Chem. 4 (2012) 621–628.
[54] C. Turro, Proc. Natl. Acad. Sci. USA 108 (2011) 17573–17574.
[55] A. Chouai, S.E. Wicke, C. Turro, J. Bacsa, K.R. Dunbar, D. Wang, R.P. Thummel,
References Inorg. Chem. 44 (2005) 5996–6003.
[56] C. Paoletti, J.B. LePecq, I.R. Lehman, J. Mol. Biol. 55 (1971) 75–100.
[1] B. Alberts, A. Johnson, J. Lewis, M. Raff, K. Roberts, P. Walter, Molecular Biology [57] Y. Liu, R. Hammitt, D.A. Lutterman, R.P. Thummel, C. Turro, Inorg. Chem. 46
of the Cell, 5th ed., Garland Science, Taylor & Francis Group, LLC, NY, 2008. (2007) 6011–6021.
[2] R. Siegel, D. Naishadham, A. Jemal, CA: Cancer J. Clin. 63 (2013) 11–30. [58] Y. Sun, S.N. Collins, L.E. Joyce, C. Turro, Inorg. Chem. 49 (2010) 4257–4262.
[3] G. Makin, Paediatr. Child Health 24 (2014) 161–165. [59] A.E. Friedman, J.C. Chambron, J.P. Sauvage, N.J. Turro, J.K. Barton, J. Am. Chem.
[4] M.J. Hannon, Chem. Soc. Rev. 36 (2007) 280–295. Soc. 112 (1990) 4960–4962.
[5] B. Armitage, Chem. Rev. 98 (1998) 1171–1200. [60] R.M. Hartshorn, J.K. Barton, J. Am. Chem. Soc. 114 (1992) 5919–5925.
[6] L.J.K. Boerner, J.M. Zaleski, Curr. Opin. Chem. Biol. 9 (2005) 135–144. [61] I. Haq, P. Lincoln, D. Suh, B. Norden, B.Z. Chowdhry, J.B. Chaires, J. Am. Chem.
[7] B. Rosenberg, Interdisciplin. Sci. Rev. 3 (1978) 134–147. Soc. 117 (1995) 4788–4796.
[8] B. Rosenberg, L. Van Camp, T. Krigas, Nature 205 (1965) 698–699. [62] Y. Liu, R. Hammitt, D.A. Lutterman, L.E. Joyce, R.P. Thummel, C. Turro, Inorg.
[9] B. Rosenberg, L. Vancamp, J.E. Trosko, V.H. Mansour, Nature 222 (1969) Chem. 48 (2008) 375–385.
385–386. [63] R. Zhao, R. Hammitt, R.P. Thummel, Y. Liu, C. Turro, R.M. Snapka, Dalton Trans.
[10] I. Kostova, Recent Patents Anti-Cancer Drug Discov. 1 (2006) 1–22. (2009) 10926–10931.
[11] M.A. Fuertes, C. Alonso, J.M. Pérez, Chem. Rev. 103 (2003) 645–662. [64] L.E. Joyce, J.D. Aguirre, A.M. Angeles-Boza, A. Chouai, P.K.L. Fu, K.R. Dunbar, C.
[12] E.R. Jamieson, S.J. Lippard, Chem. Rev. 99 (1999) 2467–2498. Turro, Inorg. Chem. 49 (2010) 5371–5376.
[13] K.S. Lovejoy, S.J. Lippard, Dalton Trans. (2009) 10651–10659. [65] J.D. Aguirre, A.M. Angeles-Boza, A. Chouai, C. Turro, J.-P. Pellois, K.R. Dunbar,
[14] R.B. Martin, Cisplatin, Verlag Helvetica Chimica Acta, Zurich, 2006, pp. Dalton Trans. (2009) 10806–10812.
181–205. [66] P.K.L. Fu, P.M. Bradley, D. van Loyen, H. Dürr, S.H. Bossmann, C. Turro, Inorg.
[15] Z.H. Siddik, Oncogene 22 (2003) 7265–7279. Chem. 41 (2002) 3808–3810.
[16] Á. Juarranz, P. Jaén, F. Sanz-Rodríguez, J. Cuevas, S. González, Clin. Transl. [67] A.B. Tossi, H. Görner, D. Schulte-Frohlinde, Photochem. Photobiol. 50 (1989)
Oncol. 10 (2008) 148–154. 585–597.
[17] I.J. Macdonald, T.J. Dougherty, J. Porphyrins Phthalocyanines 5 (2001) [68] K.L. Nguyen, M. Steryo, K. Kurbanyan, K.M. Nowitzki, S.M. Butterfield, S.R.
105–129. Ward, E.D. Stemp, J. Am. Chem. Soc. 122 (2000) 3585–3594.
[18] M.-F. Zuluaga, N. Lange, Curr. Med. Chem. 15 (2008) 1655–1673. [69] Y. Sun, L.E. Joyce, N.M. Dickson, C. Turro, Chem. Commun. 46 (2010)
[19] K. Szacilowski, W. Macyk, A. Drzewiecka-Matuszek, M. Brindell, G. Stochel, 2426–2428.
Chem. Rev. 105 (2005) 2647–2694. [70] Y. Sun, L.E. Joyce, N.M. Dickson, C. Turro, Chem. Commun. 46 (2010)
[20] R. Allison, R. Cuenca, G. Downie, P. Camnitz, B. Brodish, C. Sibata, Photodiag. 6759–6761.
Photodyn. Ther. 2 (2005) 205–222. [71] J.-P. Lecomte, A. Kirsch-De Mesmaeker, M.M. Feeney, J.M. Kelly, Inorg. Chem.
[21] R. Allison, K. Moghissi, G. Downie, K. Dixon, Photodiag. Photodyn. Ther. 8 34 (1995) 6481–6491.
(2011) 231–239. [72] L. Jacquet, J.M. Kelly, A.K.-D. Mesmaeker, J. Chem. Soc. Chem. Commun. (1995)
[22] P. Babilas, S. Schreml, M. Landthaler, R.M. Szeimies, Photodermatol. Photoim- 913–914.
munol. Photomed. 26 (2010) 118–132. [73] L. Jacquet, A. Kirsch-De Mesmaeker, J. Chem. Soc. Faraday Trans. 88 (1992)
[23] A.K. D’Cruz, M.H. Robinson, M.A. Biel, Head Neck 26 (2004) 232–240. 2471–2480.
[24] K. Moghissi, K. Dixon, M. Stringer, J. Thorpe, Photodiag. Photodyn. Ther. 6 [74] J.-P. Lecomte, A.K.-D. Mesmaeker, J.M. Kelly, A.B. Tossi, H. Görner, Photochem.
(2009) 159–166. Photobiol. 55 (1992) 681–689.
[25] A.E. O’Connor, W.M. Gallagher, A.T. Byrne, Photochem. Photobiol. 85 (2009) [75] A. Masschelein, L. Jacquet, A. Kirsch-De Mesmaeker, J. Nasielski, Inorg. Chem.
1053–1074. 29 (1990) 855–860.
[26] C. Tanielian, C. Schweitzer, R. Mechin, C. Wolff, Free Radic. Biol. Med. 30 (2001) [76] S. Monro, J. Scott, A. Chouai, R. Lincoln, R. Zong, R.P. Thummel, S.A. McFarland,
208–212. Inorg. Chem. 49 (2010) 2889–2900.
[27] G. Stochel, A. Wanat, E. Kuliś, Z. Stasicka, Coord. Chem. Rev. 171 (1998) [77] R. Lincoln, L. Kohler, S. Monro, H. Yin, M. Stephenson, R. Zong, A. Chouai,
203–220. C. Dorsey, R. Hennigar, R.P. Thummel, J. Am. Chem. Soc. 135 (2013)
[28] L. Wyld, M.W. Reed, N.J. Brown, Br. J. Cancer 77 (1998) 1621–1627. 17161–17175.
[29] N.J. Farrer, L. Salassa, P.J. Sadler, Dalton Trans. (2009) 10690–10701. [78] R.B. Nair, E.S. Teng, S.L. Kirkland, C.J. Murphy, Inorg. Chem. 37 (1998)
[30] N.A. Smith, P.J. Sadler, Philos. Trans. Roy. Soc. A: Math. Phys. Eng. Sci. 371 139–141.
(2013) 20120519. [79] T.C. Tang, H.J. Huang, Electroanalysis 11 (1999) 1185–1190.
[31] J.K. Barton, E.D. Olmon, P.A. Sontz, Coord. Chem. Rev. 255 (2011) 619–634. [80] B.S. Howerton, D.K. Heidary, E.C. Glazer, J. Am. Chem. Soc. 134 (2012)
[32] B.M. Zeglis, V.C. Pierre, J.K. Barton, Chem. Commun. (2007) 4565–4579. 8324–8327.
[33] M.G. Sauaia, R.G. de Lima, A.C. Tedesco, R.S. da Silva, J. Am. Chem. Soc. 125 [81] A.A. Brożyna, L. VanMiddlesworth, A.T. Slominski, Int. J. Cancer 123 (2008)
(2003) 14718–14719. 1448–1456.
[34] A.G. De Candia, J.P. Marcolongo, R. Etchenique, L.D. Slep, Inorg. Chem. 49 [82] A.M. Angeles-Boza, P.M. Bradley, P.K.L. Fu, M. Shatruk, M.G. Hilfiger, K.R. Dun-
(2010) 6925–6930. bar, C. Turro, Inorg. Chem. 44 (2005) 7262–7264.
[35] B. Heilman, P.K. Mascharak, Philos. Trans. Roy. Soc. A: Math. Phys. Eng. Sci. [83] A.M. Angeles-Boza, P.M. Bradley, P.K.L. Fu, S.E. Wicke, J. Bacsa, K.R. Dunbar, C.
371 (2013) 20120368. Turro, Inorg. Chem. 43 (2004) 8510–8519.
[36] U. Schatzschneider, Inorg. Chim. Acta 374 (2011) 19–23. [84] P.M. Bradley, A.M. Angeles-Boza, K.R. Dunbar, C. Turro, Inorg. Chem. 43 (2004)
[37] M. Salierno, C. Fameli, R. Etchenique, Eur. J. Inorg. Chem. 2008 (2008) 2450–2452.
1125–1128. [85] P.K.L. Fu, P.M. Bradley, C. Turro, Inorg. Chem. 40 (2001) 2476–2477.

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018
G Model
CCR-111879; No. of Pages 17 ARTICLE IN PRESS
J.D. Knoll, C. Turro / Coordination Chemistry Reviews xxx (2014) xxx–xxx 17

[86] Z. Li, S.J. Burya, C. Turro, K.R. Dunbar, Philos. Trans. Roy. Soc. A: Math. Phys. [106] F. Barragán, P. López-Senín, L. Salassa, S. Betanzos-Lara, A. Habtemariam,
Eng. Sci. 371 (2013). V. Moreno, P.J. Sadler, V. Marchán, J. Am. Chem. Soc. 133 (2011) 14098–
[87] J.L. Bear, H.B. Gray, L. Rainen, I.M. Chang, R. Howard, G. Serio, A.P. Kimball, 14108.
Cancer Chemother. Rep. Pt. 1 59 (1975) 611–620. [107] A.A. Holder, S. Swavey, K.J. Brewer, Inorg. Chem. 43 (2003) 303–308.
[88] K. Das, K.M. Kadish, J.L. Bear, Inorg. Chem. 17 (1978) 930–934. [108] A.A. Holder, D.F. Zigler, M.T. Tarrago-Trani, B. Storrie, K.J. Brewer, Inorg. Chem.
[89] R.A. Howard, A.P. Kimball, J.L. Bear, Cancer Res. 39 (1979) 2568–2573. 46 (2007) 4760–4762.
[90] K.M. Kadishi, K. Das, R. Howard, A. Dennis, J.L. Bear, Bioelectrochem. Bioenerg. [109] B. Storrie, A. Holder, K.J. Brewer, in: D. Kessel (Ed.), Society of Photo-
5 (1978) 741–753. Optical Instrumentation Engineers (SPIE) Conference Series, 2006, pp. 336–
[91] D.A. Lutterman, P.K.L. Fu, C. Turro, J. Am. Chem. Soc. 128 (2005) 738–739. 342.
[92] H.T. Chifotides, P.K.-L. Fu, K.R. Dunbar, C. Turro, Inorg. Chem. 43 (2004) [110] S. Swavey, K.J. Brewer, Inorg. Chem. 41 (2002) 6196–6198.
1175–1183. [111] S. Swavey, R.L. Williams, Z. Fang, M. Milkevitch, K.J. Brewer, in: W.B. Spillman
[93] T.N. Singh, C. Turro, Inorg. Chem. 43 (2004) 7260–7262. (Ed.), Society of Photo-Optical Instrumentation Engineers (SPIE) Conference
[94] M. Zhao, M. Clarke, J. Biol. Inorg. Chem. 4 (1999) 325–340. Series, 2001, pp. 75–83.
[95] N. Grover, N. Gupta, H.H. Thorp, J. Am. Chem. Soc. 114 (1992) 3390–3393. [112] J. Wang, S.L.H. Higgins, B.S.J. Winkel, K.J. Brewer, Chem. Commun. 47 (2011)
[96] R.N. Garner, L.E. Joyce, C. Turro, Inorg. Chem. 50 (2011) 4384–4391. 9786–9788.
[97] J. Friedrich, C. Seidel, R. Ebner, L.A. Kunz-Schughart, Nat. Protoc. 4 (2009) [113] J. Wang, J. Newman, S.L. Higgins, K.M. Brewer, B.S. Winkel, K.J. Brewer, Angew.
309–324. Chem. 125 (2013) 1300–1303.
[98] E. Wachter, D.K. Heidary, B.S. Howerton, S. Parkin, E.C. Glazer, Chem. Commun. [114] J. Wang, D.F. Zigler, N. Hurst, H. Othee, B.S.J. Winkel, K.J. Brewer, J. Inorg.
48 (2012) 9649–9651. Biochem. 116 (2012) 135–139.
[99] A.M. Palmer, B. Peña, R.B. Sears, O. Chen, M. El Ojaimi, R.P. Thummel, K.R. [115] C. Holohan, S. Van Schaeybroeck, D.B. Longley, P.G. Johnston, Nat. Rev. Cancer
Dunbar, C. Turro, Philos. Trans. Roy. Soc. A: Math. Phys. Eng. Sci. 371 (2013). 13 (2013) 714–726.
[100] D.V. Pinnick, B. Durham, Inorg. Chem. 23 (1984) 1440–1445. [116] G.A. Gentry, P.A. Morse, M.T. Dorsett, Cancer Res. 31 (1971) 909–912.
[101] C. Gaiddon, P. Jeannequin, P. Bischoff, M. Pfeffer, C. Sirlin, J.-P. Loeffler, J. [117] R.N. Garner, J.C. Gallucci, K.R. Dunbar, C. Turro, Inorg. Chem. 50 (2011)
Pharmacol. Exp. Ther. 315 (2005) 1403–1411. 9213–9215.
[102] L. Leyva, C. Sirlin, L. Rubio, C. Franco, R. Le Lagadec, J. Spencer, P. Bischoff, C. [118] M.A. Sgambellone, A. David, R.N. Garner, K.R. Dunbar, C. Turro, J. Am. Chem.
Gaiddon, J.P. Loeffler, M. Pfeffer, Eur. J. Inorg. Chem. 2007 (2007) 3055–3066. Soc. 135 (2013) 11274–11282.
[103] S. Betanzos-Lara, L. Salassa, A. Habtemariam, O. Novakova, A.M. Pizarro, [119] E. Corral, A.C. Hotze, A. Magistrato, J. Reedijk, Inorg. Chem. 46 (2007)
G.J. Clarkson, B. Liskova, V. Brabec, P.J. Sadler, Organometallics 31 (2012) 6715–6722.
3466–3479. [120] O. Novakova, J. Kasparkova, O. Vrana, P.M. van Vliet, J. Reedijk, V. Brabec,
[104] S. Betanzos-Lara, L. Salassa, A. Habtemariam, P.J. Sadler, Chem. Commun. Biochemistry 34 (1995) 12369–12378.
(2009) 6622–6624. [121] T. Respondek, R.N. Garner, M.K. Herroon, I. Podgorski, C. Turro, J.J. Kodanko, J.
[105] H. Chen, J.A. Parkinson, O. Nováková, J. Bella, F. Wang, A. Dawson, R. Gould, Am. Chem. Soc. 133 (2011) 17164–17167.
S. Parsons, V. Brabec, P.J. Sadler, Proc. Natl. Acad. Sci. USA 100 (2003) [122] C. Jedeszko, B.F. Sloane, Biol. Chem. 385 (2004) 1017–1027.
14623–14628.

Please cite this article in press as: J.D. Knoll, C. Turro, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.05.018

You might also like