Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Article

pubs.acs.org/Langmuir

Unveiling the Structure of Polytetraruthenated Nickel Porphyrin by


Raman Spectroelectrochemistry
Luís M. C. Ferreira,† Daniel Grasseschi,† Mauro S. F. Santos,† Paulo R. Martins,‡ Ivano G. R. Gutz,†
Ana Maria C. Ferreira,† Koiti Araki,*,† Henrique E. Toma,† and Lúcio Angnes*,†

Instituto de Química, Universidade de São Paulo, CEP 05508-000 Sao Paulo, SP, Brazil

Instituto de Química, Universidade Federal de Goiás, CEP 74001970 Goiania, GO, Brazil
*
S Supporting Information

ABSTRACT: The structure of polytetraruthenated nickel porphyrin was


unveiled for the first time by electrochemistry, Raman spectroelectrochemistry,
and a hydroxyl radical trapping assay. The electrocatalytic active material,
precipitated on the electrode surface after successive cycling of [NiTPyP{Ru-
(bipy)2Cl}4]4+ species in strong aqueous alkaline solution (pH 13), was found to
be a peroxo-bridged coordination polymer. The electropolymerization process
involves hydroxyl radicals (as confirmed by the characteristic set of DMPO/•OH
adduct EPR peaks) as reaction intermediates, electrocatalytically generated in the
0.80−1.10 V range, that induce the formation of Ni−O−O−Ni coordination
polymers, as evidenced by Raman spectroelectrochemistry and molecular
modeling studies. The film growth is halted above 1.10 V due to the formation
of oxygen gas bubbles.

1. INTRODUCTION mediated by electrogenerated hydroxyl radicals.3 However, the


Electrocatalytically active modified electrodes can be easily similarity in the electrochemical behavior of all of those
prepared by voltammetric cycling of nickel(II) macrocycle polymeric materials among themselves and with that of nickel
complexes such as nickel(II) tetraazamacrocycles,1−5 Ni- hydroxide29−31 is intriguing, and the elucidation of their
phtalocyanines,6−16 Ni-porphyrins,5,8,17−20 and Ni-Salen com- structure may help to shed light on the mechanism of
plexes9,11 in strongly alkaline solutions (pH ≥13). Interesting stabilization of nickel hydroxide in the alpha phase.31,32 Thus,
materials17,20−24 capable of promoting the oxidation of many the formation of nickel hydroxide after a kind of “electro-
organic compounds are generated, including alcohols, at chemically induced demetalation” of the starting Ni-TSPP
relatively low overpotentials. Malinski et al.25,26 studied the complex generating a TSPP/Ni(OH)2 conducting salt, where
electropolymerization of tetrakis(3-methoxy-4-hydroxiphenyl)- the porphyrin is present as a radical cation, has also been
porphyrin and proposed the use of carbon-fiber ultra- proposed.18 Nevertheless, the stability of such a reactive
microelectrodes modified with the resultant material as porphyrin species in a strongly alkaline medium is questionable.
amperometric sensors for the determination of nitric oxide in In fact, there are few reports showing unequivocal experimental
a single cell. The monomers were thought to be connected evidence to help elucidate their actual chemical structure, either
through the meso-methoxyphenol substituents in the nickel as hybrid materials containing Ni(OH)2 or as μ-oxo-bridged
porphyrin meso positions.20,21 However, cationic [Ni- nickel macrocycle coordination polymers. Also, no evidence of
TMPyP]4+ and anionic [Ni-TSPP]4− porphyrin complexes,20,27 possible intermediate species participating in the polymer
where TMPyP4+ = meso-tetra(pyridyl)porphyrin and TSPP4− = formation mechanism has been reported yet.3,5,8,14,33
meso-tetra(4-sulfonatephenyl)porphyrin, were also shown to Raman spectroelectrochemistry is a powerful technique for
exhibit electropolymerization properties generating similar the investigation of electrochemically generated intermediates
electroactive materials. More recently, Bedioui et al.28 prepared and products that can provide invaluable structural information
hybrid materials based on carbon nanotubes and tetrasulfo- on both the species in solution and those adsorbed on the
nated nickel(II) phthalocyanine with similar electrochemical electrode surface. In fact, pioneering work by Fleischmann,
properties. Hendra, and MacQuilan34 on the anomalous intensification of
The characteristic electrochemical behavior of polynickel Raman scattering from pyridine molecules adsorbed on silver
porphyrins and polynickel macrocycles is usually assigned to and gold electrode surfaces culminated in the discovery of
the formation and subsequent deposition of coordination
polymers of the corresponding nickel complex on the electrode Received: January 22, 2015
surface, where polymerization takes place through axial μ-oxo Revised: March 21, 2015
bridges connecting the nickel centers in a reaction that could be Published: March 26, 2015

© 2015 American Chemical Society 4351 DOI: 10.1021/acs.langmuir.5b00250


Langmuir 2015, 31, 4351−4360
Langmuir Article

Figure 1. Electronic spectrum of a 1 × 10−4 mol·L−1 [NiTPyP{Ru(bipy)2Cl}4](TFMS) solution in (a) methanol and (b) a 0.1 mol·L−1 aqueous
NaOH solution. The black and red lines correspond to the experimental and calculated spectra by deconvolution with Gaussian functions. Note that
no substitution of the chloro ligands is expected in the peripheral R = py-[Ru(bipy)2Cl]+ groups.

surface-enhanced Raman spectroscopy, SERS. Since then, 2. EXPERIMENTAL SECTION


Raman spectroeletrochemistry has been used to characterize 2.1. Chemicals and Reagents. Supermolecular complex μ-meso-
different types of interfaces, such as those based on porphyrin tetra(4-pyridyl)porphyrinatenickel(II)-tetrakis[bis(bipyridine)-
thin films obtained by interfacial polymerization35 as well as (chloro)ruthenium(II)](TFMS) 4 and sodium meso-tetra(4-
graphene and semiconductor thin films.36,37 sulfonatephenyl)porphyrinate zinc(II), hereafter denoted respectively
Tetraruthenated metalloporphyrins were extensively inves- as [NiTPyP{Ru(bipy)2Cl}4]4+ and [ZnTSPP]4−, were synthesized as
tigated as redox mediators for the oxidation of several previously reported.31,38−41 Sodium hydroxide and methanol were
purchased from Merck (Darmstadt, Germany). 5,5-Dimethyl-1-
compounds for sensor applications, exploring the [Ru- pyrroline N-oxide (DMPO) was acquired from Sigma-Aldrich Co.
(bipy)2Cl]+ peripheral complexes as electrocatalytic active and purified by distillation under reduced pressure, as recommended,
sites in slightly acidic media.38−40 Interestingly, the tetrar- before its use as a spin scavenger.42 All aqueous solutions were
uthenated nickel porphyrin, [NiTPyP{Ru(bipy)2Cl}4]4+, can prepared using ultrapure water from a Millipore Milli-Q system
also be electropolymerized in pH 13 solution by redox cycling (resistivity ≥18.2 MΩ·cm).
in the 0.0 to 1.0 V vs Ag/AgCl(KCl sat.) range, despite the 2.2. Electrode Modification and Electrochemistry. The
electrochemical characterization was performed using a μ-Autolab
presence of bulky [Ru(bipy)2Cl(pyP)]+ complexes at the meso type III potentiostat/galvanostat (EcoChemie, The Netherlands). The
positions, thus making unlikely the formation of axial μ-oxo voltammetric measurements were carried out in a conventional 10 mL
polymers. As expected, the nickel center is electrochemically cell using a (modified) glassy carbon disk working electrode
active only in highly alkaline media, where the modified (geometric area: 0.071 cm2), Ag/AgCl(KCl sat.) reference electrode,
electrode exhibits a very sharp couple of waves at around 0.4− and a platinum wire as an auxiliary electrode. Electrodes were modified
0.5 V, which are assigned to a Ni3+/Ni2+ process. Furthermore, as previously reported 19 by performing 5 or 50 successive
voltammetric cycles in a 1.0 × 10−4 mol·L−1 solution of [NiTPyP-
the peripheral ruthenium complexes seem to play an important {Ru(bipy)2Cl}4]4+ species in 0.1 mol·L−1 NaOH supporting electro-
role in film growth because the electropolymerization process lyte in the 0.00 to 0.90 V range (scan rate = 0.1 V·s−1). The modified
was shown to be remarkably dependent on the limiting anodic electrodes were washed with the electrolyte solution and reserved for
potential used in the redox cycling,19 providing new strategies further use. All potentials are referred to the Ag/AgCl(KCl sat.) reference
to investigate further the nature of that polymeric material. electrode.
Accordingly, the influence of hydroxide ions, the role played by The experimental conditions for electrode modification and
electrochemical assays were chosen after careful study by varying the
Ru and Ni sites in the electropolymerization process and film deposition pH and limiting the positive potential as shown respectively
growth, and the composition and structure of poly-[NiTPyP- in Figures S4 and S5. The polymeric material was not formed at pH 14
{Ru(bipy)2Cl}4] were elucidated by a detailed electronic and probably because of the shift in the oxygen gas evolution reaction to
Raman spectroelectrochemical study. lower potentials.

4352 DOI: 10.1021/acs.langmuir.5b00250


Langmuir 2015, 31, 4351−4360
Langmuir Article

2.3. Electronic Spectra. Electronic spectra in solution were bipy ligands. The Ru(II) ligand field transitions were found at
recorded in an HP 8453 diode array spectrophotometer using a 10.0 355 nm whereas the bands at 434 and 486 nm were assigned to
mm optical path quartz cuvette. The electronic reflectance spectra of Ru(II) → bipy and dπ → pπ* metal-to-ligand charge-transfer
modified platinum electrodes were obtained using an Analytical transitions (MLCT). The changes observed in the MLCT
Spectral Devices (ASD) Field Spec fiber optics probe spectropho-
tometer equipped with a tungsten lamp. Platinum electrodes were
bands in methanol and NaOH aqueous solution can be
modified following the procedure previously described for the attributed to solvatochromic effects. The characteristic
modification of glassy carbon (GC) electrodes. The spectrum was porphyrin ring bands at 408, 526, and 558 nm were assigned
registered in reflectance mode and converted to the absorbance scale to the Soret, β, and α transitions, respectively.
using the ASD Field Spec equipment software. According to the four frontier orbitals model48 for the
2.4. Raman Spectra and Spectroelectrochemistry. Raman porphyrin electronic states, the Soret band is marked by a
spectra were recorded using a WITec Alpha 300R confocal Raman transition from the HOMO (with a2u symmetry) to the LUMO
microscope equipped with an EM-CCD analyzer and Ar (488 nm), (eg) orbital. The a2u orbital exhibits a lower electronic density
Nd:YAG (532 nm), and He−Ne (633 nm) lasers. The Raman on the pyrrolic carbons, being more sensitive to the nature of
spectrum of solid [NiTPyP{Ru(bipy)2Cl}4](TFMS)4 was recorded at
488 and 532 nm (power = 0.15 and 2.00 mW) using an integration
substituents at the meso positions and the coordinated metal
time of 10 s. The Raman spectrum of a 1.0 × 10−3 mol·L−1 solution of ion and respective axial ligands. Therefore, changes in the
this supramolecular complex in aqueous 0.1 mol·L−1 NaOH was electronic properties of the peripheral ruthenium complexes
recorded at 488 and 532 nm (power = 7.50 mW for both lasers) using and the coordination of hydroxide anions to the nickel
an integration time of 120 s. porphyrin axial positions should cause a blue shift and decrease
Raman spectroelectrochemistry experiments were carried out using in the intensity of the Soret band, as observed in Figure 1b.
a homemade three-electrode spectroelectrochemical cell (Figure S1 in Accordingly, one can assume that [Ni(OH)2TPyP{Ru-
Supporting Information) and a PalmSens potentiostat/galvanostat. (bipy)2Cl}4]2+ is the predominant species in the 0.1 mol·L−1
The laser was focused on the GC electrode surface and three NaOH aqueous solution.
successive Raman spectra (λ = 532 nm, power = 0.50 mW, integration
time = 10 s) recorded after surface modification with a poly-
After 50 voltammetric cycles in the 0.00 to 0.90 V range in
[NiTPyP{Ru(bipy)2Cl}4] film while applying a potential in the 0.05 to aqueous NaOH solution (0.1 mol·L−1), a thin film of poly-
0.85 V range. [NiTPyP{Ru(bipy)2Cl}4] is deposited on the platinum disk
2.5. Electron Paramagnetic Resonance Spectroscopy. The electrode surface. This exhibits red-shifted porphyrin transitions
hypothesis of a radical species being the key intermediate species for (Soret, β and α) consistent with formation of a π-stacked
the formation of the poly-[NiTPyP{Ru(bipy)2Cl}4] film was porphyrin material (Figure 2), whereas the MLCT bands of the
investigated using a GC electrode modified with an electrostatically peripheral ruthenium complexes exhibited only a slight red shift
assembled [NiTPyP{Ru(bipy)2Cl}4]4+/[ZnTSPP]4− film39 polarized and broadening indicating a much lower degree of
at 0.90 V. The spin-trapping reaction for hydroxyl radical detection intermolecular interactions.
was carried out in a specially designed electrochemical flow cell
(Figure S2 in Supporting Information). A stream of a 0.1 mol·L−1
NaOH solution was continuously passed over the modified GC surface
and merged with a 2.0 × 10−4 mol·L−1 DMPO solution at a confluence
point placed immediately after the working electrode. A peristaltic
pump was used to propel both solutions at 0.6 mL·min−1. The
resultant effluent solution was collected in an appropriate flask and
injected into a 200 μL flat quartz cell placed in the cavity of a Bruker
EMX EPR spectrometer operating at the X band (9.5 GHz). Spectra
were registered at room temperature using the following parameters:
modulation frequency = 100 kHz, modulation amplitude = 1 G, and
microwave power = 20.21 mW.
2.6. Molecular Modeling. The ground-state geometry optimiza-
tion and the Raman spectra calculation for the target species were
carried out using the Gaussian 09W software employing density
functional theory (DFT)43,44 and the B3LYP hybrid functional45
(Becke’s gradient-corrected exchange correlation46 in conjunction with
the Lee−Yang−Parr correlation functional with three parameters47).
The 6-311G++(d,p) basis set for C, N, and H atoms and the
LANL2DZ basis set for the nickel atom, with their respective
pseudopotentials for the inner shell, were employed in the calculations. Figure 2. Electronic spectra of poly-[NiTPyP{Ru(bipy)2Cl}4]
The assignment of the vibrational spectra was carried out by deposited on a platinum disk electrode surface after successive cycling
comparison with literature data using GaussView 05 and Gabedit (50 times) in a pH 13 aqueous 1.0 × 10−4 mol·L−1 [NiTPyP{Ru-
2.4.0 software. (bipy)2Cl}4]4+ solution in the 0.00 to 0.90 V range (black line).
Calculated absorption spectrum (red line) and set of bands obtained
by deconvolution of the actual spectrum using Gaussian functions.
3. RESULTS AND DISCUSSION
3.1. Electronic Spectroscopy. The electronic spectrum of 3.2. Electrochemical Behavior and Characterization.
[NiTPyP{Ru(bipy)2Cl}4](TFMS)4 in methanol solution (Fig- The cyclic voltammograms (CVs) corresponding to 50
ure 1a) shows absorption bands at 209, 244, 291, 355, 408, 434, successive redox cycles of a platinum disk electrode in a pH
486, 526, and 558 nm, consistent with the presence of two 13 aqueous 1.0 × 10−4 mol·L−1 [NiTPyP{Ru(bipy)2Cl}4]-
chromophores, the nickel(II) porphyrin and the peripheral (TFMS)4 solution, in the 0.0−0.9 V range, are presented in
ruthenium complexes.41 The bands at 291, 355, 434, and 486 Figure 3a. The precipitation of poly-[NiTPyP{Ru(bipy)2Cl}4]
nm were assigned to the Ru complexes, where the 291 nm band on the electrode surface is confirmed by the rise of a couple of
was assigned to characteristic π → π* internal transitions of the sharp anodic and cathodic peaks respectively at 0.49 and 0.40
4353 DOI: 10.1021/acs.langmuir.5b00250
Langmuir 2015, 31, 4351−4360
Langmuir Article

and allows us to explore only the electrochemical properties of


the peripheral ruthenium complexes because the tetraanionic
porphyrin stacks on the top of the tetracationic porphyrin
species acting as a spacer, thus sterically hindering the
nickel(II) axial positions and inhibiting the formation of μ-
oxo or μ-peroxo bridges that may consume that radical species.
Interestingly, the EPR spectrum of the resulting solution
(Figure 4) clearly exhibited the typical signals of the

Figure 3. (a) Successive 50 cyclic voltammograms recorded while


monitoring the growth of poly-[NiTPyP{Ru(bipy)2Cl}4] on a glassy
carbon electrode from an aqueous 1.0 × 10−4 mol·L−1 monomer
solution in 0.1 mol·L−1 NaOH solution (pH 13) at 0.100 V·s−1 and Figure 4. EPR spectrum of the solution generated in the electro-
(b) CVs corresponding to the first three cycles in the 0.0 to 0.9 V chemical flow cell set with a GC electrode modified with a
range where the 0.0 to 0.7 V region is shown in detail in the inset. CV [NiTPyP{Ru(bipy)2Cl}4]4+/[ZnTPPS]4− film and polarized at 0.90
in pure electrolyte solution (black line), first (red line), second (blue V, showing the characteristic signals of the DMPO/•OH spin adduct.
line), and third (green line) scan in [NiTPyP{Ru(bipy)2Cl}4]-
(TFMS)4 solution. DMPO/•OH adduct, with a hyperfine constant of 15.6 G,
consistent with the reference value of 15.01 G.50−53 This result
V, assigned to the Ni3+/Ni2+ redox couple. In particular, the indicates the formation of significant numbers of hydroxyl
first three cycles (Figure 3b) were found to be important in the radicals at the GC electrode modified with electrostatically
investigation of the mechanism of polymer formation. assembled porphyrins and polarized at 0.9 V.
The rise of an irreversible anodic peak at around 0.80 V, The oxidation of water to molecular oxygen by the
starting from the very first cycle in the presence of the [RuIII(bipy)3] complex in alkaline media was previously
tetraruthenated nickel porphyrin, is evident in all 50 voltammo- reported in the literature.54−58 A similar reaction probably
grams, suggesting the possible contribution of the peripheral can be promoted by the [RuIII(bipy)2Cl(pyP)] species under
ruthenium complexes to the mechanism of polymer formation. appropriated electrochemical conditions. Nevertheless, because
Ru3+/Ru2+ is a reversible monoelectronic process38,39,49 such there was no gas evolution at the modified electrode polarized
that the intense irreversible wave can be assigned to an at 0.90 V, that reaction should not be significant under our
electrocatalytic process. In fact, the electrolyte discharge is experimental conditions. Instead, in a 0.1 mol·L−1 NaOH
anticipated in 0.4 V (Figure S3 in Supporting Information) solution, hydroxyl radicals seem to be generated as the major
when the tetraruthenated nickel porphyrin is present in oxidation product, suggesting that it may be the key
solution, suggesting the key role of the peripheral ruthenium intermediate species involved in the electropolymerization
complexes in that electrochemical reaction. Considering the mechanism leading to the formation of poly-[NiTPyP{Ru-
absence of any other species except those present in the (bipy)2Cl}4].
supporting electrolyte solution, it should involve the electro- Thus, the [RuIII(bipy)2Cl(pyP)] species generated in alkaline
catalytic oxidation of hydroxide anions to hydrogen peroxide media at potentials above 0.80 V should be the active species
anion, molecular oxygen, or hydroxyl radical. The last responsible for the formation of hydroxyl radicals, either by
hypothesis is particularly interesting because it may be involved direct electron transfer or involving an intermediate species as
in the electropolymerization mechanism18 leading to the proposed by Ledney et al.57 (Scheme 1). Direct electron
formation of poly-[NiTPyP{Ru(bipy)2Cl}4] species. transfer is thermodynamically forbidden (OH• + e− = OH, E0 =
The electrochemical generation of dioxygen on the electrode 2.02 V vs SHE), but a reaction mechanism involving the
surface at sufficiently positive potentials is well documented in nucleophilic attack of OH− or H2O molecules on the para
the literature, but the formation of the other two species is position of one of the bipy ligands produced a protonated
much more scarce. Thus, the hypothesis of hydroxyl radical Ru(III) adduct whose rapid deprotonation reaction leads to the
formation mediated by peripheral Ru(III) complexes was redistribution of the charge density corresponding to the pair of
pursued by EPR. For this purpose, an especially designed flow electrons and the formation of a hydroxylated bipyridyl radical.
cell was constructed in order to trap the •OH species with The deprotonation of the bipy N−H intermediate is facilitated
DMPO soon after being generated on a GC electrode modified by the high hydroxide concentration in the electrolyte such that
with a thin film of electrostatically assembled38 [NiTPyP{Ru- the subsequent reduction of the peripheral complex to the
(bipy)2Cl}4]4+/[ZnTSPP]4− material. This is quite insoluble Ru(II) state leads to the homolysis of the C−OH bond,
4354 DOI: 10.1021/acs.langmuir.5b00250
Langmuir 2015, 31, 4351−4360
Langmuir Article

Scheme 1. Possible Mechanism for the Electrocatalytic Generation of Hydroxyl Radicals Mediated by Peripheral
[RuIII(bipy)Cl(pyP)] Complexes in Strong Aqueous Alkaline Solutiona

a
Adapted from Ledney et al.57

Figure 5. Normalized Raman spectra of [NiTPyP{Ru(bipy)2Cl}4](TFMS)4 powder (red line) and the [Ni(OH)2TPyP{Ru(bipy)2Cl}4]2+ complex
present in aqueous 0.1 mol·L−1 NaOH solution (black line), acquired using the (a) 488 and (b) 532 nm laser lines.

releasing a •OH radical species and regenerating the bipy (ECE) mechanism taking place in a narrow potential range
ligand. Then, the complex is rapidly oxidized back to the (0.80 V ≤ E < 1.10 V) and involving electrocatalytically
Ru(III) state by the electrode, starting a new electrocatalytic generated hydroxyl radicals. These highly reactive species
cycle. should trigger some sort of polymerization process leading to
The formation of poly-[NiTPyP{Ru(bipy)2Cl}4] is strongly the deposition of a slightly soluble material as a thin film on the
dependent on that electrochemical process because the electrode surface. However, no clear evidence of what is
characteristic Ni3+/Ni2+ waves at 0.50 and 0.40 V (peaks III responsible for the polymerization process has been reported.
and IV in Figure 3b) are not observed when the potential Accordingly, Raman spectroscopy and spectroelectrochemistry
sweep is limited to values lower than 0.80 V, and there is no studies were carried out to shed light on the structure of such a
significant production of the [RuIII(bipy)2(pyP)Cl] complex. polymeric material.
Also, potentials higher than 1.10 V promote the formation of 3.3. Raman Spectroscopy and Spectroelectrochemis-
molecular oxygen, and bubbles released at the modified try. Raman studies were carried out in order to characterize the
electrode surface inhibit the deposition of that polymeric poly-[NiTPyP{Ru(bipy)2Cl}4] deposited on a glassy carbon
material. Thus, the above-described mechanism seems to electrode surface by comparing the spectra of [NiTPyP{Ru-
explain why the voltammetric pattern in alkaline media is (bipy)2Cl}4](TFMS)4 in the solid state, in an aqueous 0.1 mol·
different from that obtained in acidic media39 as well as the L−1 NaOH solution (Figure 5), and as a polymeric film after 5
dependence of the film growth on pH and the limiting anodic and 50 voltammetric redox cycles in the 0.0 to 0.9 V range
potential. The intensity of the reversible redox couple with Epa (Figure 6) at a scan rate of 0.10 V·s−1.
= 0.31 and Epc = 0.24 V (peaks I and II in Figure 3B) does not The characteristic porphyrin and bipyridine CC and C
change significantly as a function of the number of scans and N stretching modes can be seen in the 1450−1650 cm−1 range
can be assigned to a soluble species present in the electrolyte using the 532 and 488 nm laser lines. The 532 nm laser is in
solution. resonance with the nickel porphyrin β transition so that the
In summary, the formation of poly-[NiTPyP{Ru(bipy)2Cl}4] intensification of the corresponding normal modes can be
proceeds via an electrochemical−chemical−electrochemical explained by the Herzberg−Teller mechanism,59 where the
4355 DOI: 10.1021/acs.langmuir.5b00250
Langmuir 2015, 31, 4351−4360
Langmuir Article

Table 1. Tentative Assignment of the Raman Spectrum of


the [NiTPyP{Ru(bipy)2Cl}4]4+ Species in Aqueous 0.1 mol·
L−1 NaOH Solution in Comparison to the Spectrum of the
Pure [NiTPyP{Ru(bipy)2Cl}4](TFMS)4 Powder Registered
Using the 488 nm Laser Line and Typical Vibrational Modes
of the Peripheral [Ru(bipy)2Cl]+ Complex60−62
theoret. exp: 488 nm (in
[NiP] Ru(bipy)2Cl [NiTRP] NaOH solution) tentative assignment
3133 νsiCβ−H
3117 νassCβ−H
3087
1632 1600 1609 1608 νassCmCα
νsi CC and νsi CN
(bipy)
1555 1558 1557 1565 νass CC and νass CN
(bipy)
νsiCmCα
Figure 6. Theoretical Raman spectra of the [NiP(O2)]1− species νsiCβCβ
(top), where P represents the porphyrin ring. Raman spectra of the 1493 1487 1491 1490 νassCβCβ
poly-[NiTPyP{Ru(bipy)2Cl}4] species on a glassy carbon electrode νassCC and
after 50 (middle) and 5 (bottom) voltammetric cycles in the 0.00 to νassCN(bipy)
0.90 V range, in 0.1 mol·L1 aqueous NaOH solution. 1447 1455 1438 νsiCβCβ + νsiCmCα
(out of phase)
1369 1379
1348 1352 1355 δsi,opNCC +
asymmetric modes should be more enhanced than the δsi,ipCαCmCα (out
of phase)
symmetric ones. Accordingly, the asymmetric CC and C
νNiN
N stretching modes at 1515 cm−1 and the asymmetric
1338 1321 1324 νassNC
porphyrin ring deformation modes at 454 and 116 cm−1 are
1319
the most enhanced ones.
1274 1271 ν(bipy)
The spectral profile and peak intensities of [NiTPyP{Ru-
1259 1252 νCm−py + δipNi−
(bipy)2Cl}4]4+ species in aqueous 0.1 mol·L−1 NaOH solution OH
closely resemble those obtained for the pure powder as 1217 1217 ν(py)
described above, confirming that no degradation takes place 1168 1170 1173 1177 δipCβH
when the tetraruthenated nickel porphyrin is dissolved in 1090 1094
strongly alkaline solution (pH 13). The intensity of the ν(Ru− 1052 1064 1070 1066 δsi,ipNNiN
Cl) stretching mode at 253 cm−1 (Table 1) decreased in NaOH 974 1027 1024 1021 ring breathing (P +
solution when probed with both 488 and 532 nm laser lines. bipy) + νsiNiN
This suggests that hydroxide anions may be substituting for the 1004 989
chloride ligand in the Ru coordination sphere, but no evidence 889 901
of such a reaction could be found electrochemically. The bands 843
assigned to the Ni−O stretching and Ni−O−H angle 807 δsi,ipCαCβCβ +
δsi,ipCαCmCα
deformation modes were found respectively at 405 and 1252
cm−1, indicating that hydroxo ligands are coordinated to the Ni 719 744
696
porphyrin axial positions. Accordingly, one can assume that
650 662 669 668 δsi,opCαCβH +
[Ni(OH)2TPyP{Ru(bipy)2Cl}4]2+ is the major species present δsi,opCH(py)
in aqueous 0.1 mol·L−1 aqueous NaOH solution, but the 601
presence of minor amounts of supramolecular nickel porphyrin 557
containing [Ru(bipy)2(pyP)(OH)]+ groups cannot be com- 447 459
pletely ruled out. The conventional electrochemical behavior of 409 405 νNiO + δsi,opP
a [NiTPyP{Ru(bipy)2Cl}4]4+ solution was recovered after the 389 381
pH was adjusted to 5. 371 371 370 νRuN
The tentative assignment of the Raman spectra shown in 348 320 325 νsiNiN
Table 1 was carried out in comparison to spectra of the 253 νRuCl
porphyrin and its derivatives60−62 found in the literature, in 216 230 214 νassNiN + δassNiN
addition to the theoretical Raman spectra of [NiP], [NiP- 116 δsi,opP + δsi,opNiN
(OH)]−1, and [NiP(O2)]−1 species, where P represents the a
[NiTRP] = [NiTPyP{Ru(bipy)2Cl}4]4+ complex; [NiP] = non-
porphyrin ring. These species were considered in the substituted Ni porphyrin.
theoretical calculations assuming the possibility that the
deposited material could be a coordination polymer where
nickel porphyrin units are connected through hydroxide or Poly-[NiTPyP{Ru(bipy)2Cl}4] was deposited as a thin film
peroxide bridges. The optimized structures and respective from an aqueous NaOH solution such that hydroxide anions
theoretical Raman spectra are shown in the Supporting are coordinated to the axial positions of the nickel(II)
Information (Figures S6 and S7). porphyrin. The electrochemical and Raman data suggest the
4356 DOI: 10.1021/acs.langmuir.5b00250
Langmuir 2015, 31, 4351−4360
Langmuir Article

formation of μ-hydroxo, μ-peroxo, or μ-oxo bridges connecting


adjacent nickel porphyrin units. And according to the radical
trapping assay, hydroxyl radicals should be involved in the
electropolymerization mechanism.
Thus, the formation of poly-[NiTPyP{Ru(bipy)2Cl}4] was
monitored by Raman spectroelectrochemistry in order to prove
that hypothesis and identify the most probable bridging group.
The spectrum of the material deposited on the GC electrode
surface after 5 and 50 voltammetric cycles is shown in Figure 6,
in comparison to the spectrum of [NiTPyP{Ru(bipy)2Cl}4]4+
species in NaOH solution, where it is apparent that the rise of
an intense band at 399 cm−1 is assigned to a Ni−O stretching
mode.
The O−O stretching generally appears as high-intensity
bands in the 800−1300 cm−1 range depending on its
oxidization state and bond order.63,64 According to the
theoretical Raman spectra of the [NiP(O2)]1− species (Figure
6, top), the O−O stretching mode should be found around
1370 cm−1 and the O−O bond length should be 1.31 Å, as
expected for the formation of a peroxo bridge with high charge
delocalization into the porphyrin ring. Comparing the
theoretical spectra with that of poly-[NiTPyP{Ru(bipy)2Cl}4]
obtained after 5 and 50 voltammetric cycles, one can see the
rise of new bands at 1382 and 1236 cm−1, which can be
assigned to a ν(O−O) mode with slightly higher bond order
than that theoretically predicted for a peroxide. Furthermore,
the bands at 249 and 209 cm−1 can be assigned to δ(NiOO)
angular deformation modes, reinforcing the hypothesis of a μ-
peroxo-bridged Ni porphyrin polymer. No additional bands
assignable to μ-hydroxo or μ-oxo bridged species could be
found in the characteristic 800 to 1300 cm−1 range.
A red shift and broadening of most of the Raman peaks were
observed in the polymeric material obtained after 50 cycles as
compared to 5 cycles, indicating that intermolecular inter-
actions become stronger as the electropolymerization process
proceeds in NaOH solution. This behavior can be rationalized
by a higher degree of π-stacking of nickel porphyrin rings due
to the formation of a more densely packed phase, as expected
for the film growth as a function of the number of voltammetric
cycles.
The negative charge density on the O−O bridge is
dependent on the charge delocalization to the nickel porphyrin
ring, thus changing its apparent oxidation state. For example,
the band at 810 cm−1 can be assigned to a ν(Ni3+−O)
vibrational mode, which presents a higher frequency than the
analogous ν(Ni2+−O) mode at 405 cm−1. The Ru4+O
stretching mode also can be found around 800 cm−1; however,
this was disregarded because the possibility of chloride ligand Figure 7. Raman spectroelectrochemistry of poly-[NiTPyP{Ru-
substitution by hydroxide was shown to be very low, according (bipy)2Cl}4] on a glassy carbon electrode in 0.1 mol·L−1 aqueous
to our electrochemical studies as a function of the pH of the NaOH solution in three complementary wavenumber ranges.
electrolyte solution.
Finally, poly-[NiTPyP{Ru(bipy)2Cl}4] was studied by formation of Ni−O−O−Ni peroxo bridges, involving hydroxyl
Raman spectroelectrochemistry (Figure 7) in the 0.00 to 0.70 radicals as an intermediate. This is confirmed by the rise of the
V range, making it possible to correlate the structural changes 1234 and 209 cm−1 bands assigned to the ν(O−O) and
associated with a given redox process and reinforce the δ(NiOO) modes, respectively. Additionally, the red shift of the
evidence supporting the proposed structure. The first point to porphyrin ν(CC) and ν(CN) modes at 1610 and 1560 cm−1
be highlighted here is the absence of a typical Ni(OH)2 band shows the π stacking of the nickel porphyrin rings as the
around 475 cm−1 considering the whole potential range (Figure electropolymerization process proceeds. These vibrational
7a−c), thus ruling out the hypothesis of electrochemically modes were not significantly influenced when higher potentials
induced decomposition of the NiP complex65 and the were applied, confirming the assignment of the anodic wave at
formation of authentic Ni(OH)2. The spectrum of the material 0.52 V to the Ni2+/Ni3+ process. This conclusion is reinforced
formed after five voltammetric cycles (Figure 5a) showed that by the saturation behavior of the intensity of the 810 cm−1
the oxidation of Ni2+ to Ni3+ around 0.50 V is coupled with the peak, assigned to the ν(Ni3+−O) mode, that reaches its
4357 DOI: 10.1021/acs.langmuir.5b00250
Langmuir 2015, 31, 4351−4360
Langmuir Article

maximum in the 0.45 to 0.55 V range. In summary, the Raman


spectroelectrochemistry studies confirmed the hypothesis that
■ REFERENCES
(1) Azad, U. P.; Ganesan, V. Influence of Metal Nanoparticles on the
poly-[NiTPyP{Ru(bipy)2Cl}4] is constituted by stacked nickel Electrocatalytic Oxidation of Glucose by Poly(NiIIteta) Modified
macrocycles connected by μ-peroxo bridges instead of hybrid Electrodes. Electroanalysis 2010, 22, 575−583.
composites of authentic Ni(OH)2 formed upon electrochemi- (2) Cataldi, T. R. I.; Centonze, D.; Ricciardi, G. Electrode
cally induced decomposition of the nickel porphyrin complex. Modification with a Poly(NiII -tetramethyldibenzo-tetraaza[14]-
This conclusion is also supported by the fact that the CV and annulene) film. Electrochemical Behavior and Redox Catalysis in
absorption spectral profile of the starting [NiTPyP{Ru- Alkaline Solutions. Electroanalysis 1995, 7, 312−318.
(bipy)2Cl}4]4+ complex is recovered upon neutralization of (3) Roslonek, G.; Taraszewska, J. Electrocatalytic oxidation of
the polymeric material. alcohols on glassy carbon electrodes electrochemically modified with
nickel tetraazamacrocyclic complexes: mechanism of film formation. J.
Electroanal. Chem. 1992, 325, 285−300.
4. CONCLUSIONS (4) Taraszewska, J.; Roslonek, G. Voltammetric behaviour of Ni(II)
Spectroscopic and Raman spectroelectrochemical studies were and Co(III) tetraazamacrocyclic complexes containing a pendant
amino group. J. Electroanal. Chem. 1991, 297, 245−255.
carried out, demonstrating for the first time that the polymeric
(5) Trevin, S.; Bedioui, F.; Villegas, M. G. G.; Bied-Charreton, C.
material, electrochemically deposited from strongly alkaline Electropolymerized nickel macrocyclic complex-based films: design
solutions of the [NiTPyP{Ru(bipy)2Cl}4]4+ complex, is a μ- and electrocatalytic application. J. Mater. Chem. 1997, 7, 923−928.
peroxo-bridged coordination polymer. The electropolymeriza- (6) Agboola, B.; Nyokong, T. Electrocatalytic oxidation of
tion process involves the reaction of hydroxyl radicals, as chlorophenols by electropolymerised nickel(II) tetrakis benzylmer-
confirmed by EPR using DMPO as a spin-trapping agent, with capto and dodecylmercapto metallophthalocyanines complexes on
the hydroxide anions coordinated to the nickel porphyrin axial gold electrodes. Electrochim. Acta 2007, 52, 5039−5045.
positions. In fact, the oxidation of the peripheral ruthenium (7) Altamar, L.; Fernández, L.; Borras, C.; Mostany, J.; Carrero, H.;
complexes to Ru(III) leads to the electrocatalytic formation of Scharifker, B. Electroreduction of chloroacetic acids (mono-, di- and
hydroxyl radicals in the 0.80−1.10 V range, inducing the tri-) at poly-Ni(II)-tetrasulfonated phthalocyanine gold modified
electrode. Sens. Actuators, B 2010, 146, 103−110.
formation of Ni−O−O−Ni bridges. However, the film growth
(8) Berríos, C.; Marco, J. F.; Gutiérrez, C.; Ureta-Zañartu, M. S.
is halted above 1.10 V because of the formation of oxygen gas Study by XPS and UV-Visible and DRIFT Spectroscopies of
bubbles that disturb the electrode surface, precluding the Electropolymerized Films of Substituted Ni(II)-p-Phenylporphyrins
adhesion of poly-[NiTPyP{Ru(bipy)2Cl}4]. The presence of μ- and -Phthalocyanines. J. Phys. Chem. B 2008, 112, 12644−12649.
peroxo bridges was confirmed by the rise of Raman peaks (9) Berríos, C.; Ureta-Zañartu, M. S.; Gutiérrez, C. Impedance study
assigned to ν(O−O) and δ(NiOO) at 1234 and 206 cm−1 and of electropolymerized films of polyNi(II)-macrocycles. Electrochim.
the enhancement of the 810 cm−1 peak assigned to the ν(Ni3+− Acta 2007, 53, 792−802.
O) vibrational mode as Ni2+(O2−2) is oxidized to the respective (10) Chauke, V.; Matemadombo, F.; Nyokong, T. Remarkable
Ni3+ species. Our findings probably can be extended to other sensitivity for detection of bisphenol A on a gold electrode modified
nickel porphyrins. with nickel tetraamino phthalocyanine containing Ni−O−Ni bridges.


J. Hazard. Mater. 2010, 178, 180−186.
(11) Cortez, L.; Berríos, C.; Yáñez, M.; Cárdenas-Jirón, G. I.
ASSOCIATED CONTENT Theoretical study of the binding nature of glassy carbon with
*
S Supporting Information
nickel(II) phthalocyanine complexes. Chem. Phys. 2009, 365, 164−
169.
Scheme of Raman spectroelectrochemical cell. CV of GCE in (12) Khene, S.; Lobb, K.; Nyokong, T. Interaction between nickel
NaOH in the presence and absence of nickel porphyrin. hydroxy phthalocyanine derivatives with p-chlorophenol: Linking
Successive CVs as a function of pH. CVs as a function of anodic electrochemistry experiments with theory. Electrochim. Acta 2010, 56,
limiting potential in NaOH. DFT-optimized structures and 706−716.
calculated DFT Raman spectra of NiP, [NiP(OH)]−1, and (13) Mugadza, T.; Nyokong, T. Facile electrocatalytic oxidation of
[NiP(O2)]−1. This material is available free of charge via the diuron on polymerized nickel hydroxo tetraamino-phthalocyanine
Internet at http://pubs.acs.org. modified glassy carbon electrodes. Talanta 2010, 81, 1373−1379.


(14) Obirai, J.; Bedioui, F.; Nyokong, T. Electro-oxidation of phenol
and its derivates on poly-Ni(OH)TPhPyPc modified vitreous carbon
AUTHOR INFORMATION electrodes. J. Electroanal. Chem. 2005, 576, 323−332.
Corresponding Authors (15) Pontié, M.; Gobin, C.; Pauporté, T.; Bedioui, F.; Devynck, J.
Electrochemical nitric oxide microsensors: sensitivity and selectivi-
*E-mail: koiaraki@iq.usp.br. Phone: ++ 55 11 3091 8513. tyvcharacterisation. Anal. Chim. Acta 2000, 441, 175−185.
*E-mail: luangnes@iq.usp.br. Phone: ++ 55 11 3091 3828. Fax: (16) Ureta-Zañartu, M. S.; Alarcón, A.; Muñoz, G.; Gutiérrez, C.
++ 55 11 3091 3781. Electrooxidation of methanol and ethylene glycol on gold and on gold
Notes modified with an electrodeposited polyNiTSPc film. Electrochim. Acta
The authors declare no competing financial interest. 2007, 52, 7857−7864.


(17) Ciszewski, A.; Milczarek, G. Electrocatalytic oxidation of
alcohols on glassy carbon electrodes electrochemically modified by
ACKNOWLEDGMENTS conductive polymeric nickel(II) tetrakis(3-methoxy-4-hydroxyphenyl)
We thank Fundaçaõ de Amparo à Pesquisa do Estado de São porphyrin film. J. Electroanal. Chem. 1996, 413, 137−142.
(18) Pérez-Morales, M.; Muñ oz, E.; Martín-Romero, M. T.;
Paulo (FAPESP), Conselho Nacional de Desenvolvimento
́ Camacho, L. Anodic Electrodeposition of NiTSPP from Aqueous
Cientifico e Tecnológico (CNPq), and Coordenaçaõ de Basic Media. Langmuir 2005, 21, 5468−5474.
Aperfeiçoamento de Pessoal de Ensino Superior (CAPES) for (19) Quintino, M. d. S. M.; Winnischofer, H.; Nakamura, M.; Araki,
financial support and a fellowship for D.G. (FAPESP 2011/ K.; Toma, H. E.; Angnes, L. Amperometric sensor for glucose based
00037-6). We also thank Dr. Tiago Araujo Matias for helpful on electrochemically polymerized tetraruthenated nickel-porphyrin.
discussions. Anal. Chim. Acta 2005, 539, 215−222.

4358 DOI: 10.1021/acs.langmuir.5b00250


Langmuir 2015, 31, 4351−4360
Langmuir Article

(20) Trévin, S.; Bedioui, F.; Devynck, J. New electropolymerized based on tetraruthenated porphyrin modified electrodes. Anal. Chim.
nickel porphyrin films. Application to the detection of nitric oxide in Acta 1996, 329, 91−96.
aqueous solution. J. Electroanal. Chem. 1996, 408, 261−265. (39) Azevedo, C. M. N.; Araki, K.; Angnes, L.; Toma, H. E.
(21) Ciszewski, A.; Milczarek, G. Glassy carbon electrode modified Electrostatically Assembled Films for Improving the Properties of
by conductive, polymeric nickel(II) porphyrin complex as a 3D Tetraruthenated Porphyrin Modified Electrodes. Electroanalysis 1998,
homogeneous catalytic system for methanol oxidation in basic media. 10, 467−471.
J. Electroanal. Chem. 1997, 426, 125−130. (40) Martins, P. R.; Popolim, W. D.; Nagato, L. A. F.; Takemoto, E.;
(22) Fish, J. R.; Kubaszewski, E.; Pea, A.; Malinski, T.; Kaczor, J.; Araki, K.; Toma, H. E.; Angnes, L.; Penteado, M. D. V. C. Fast and
Kus, P.; Czuchajowski, L. Synthesis and Electrochemistry of reliable analyses of sulphite in fruit juices using a supramolecular
Conductive Copolymeric Porphyrins. Chem. Mater. 1992, 4, 795−803. amperometric detector encompassing in flow gas diffusion unit. Food
(23) Hrbác,̌ J.; Gregor, Č .; Machová, M.; Králová, J.; Bystroň, T.; Č íž, Chem. 2011, 127, 249−255.
M.; Lojek, A. Nitric oxide sensor based on carbon fiber covered with (41) Toma, H. E.; Araki, K. Spectroelectrochemistry And Electro-
nickel porphyrin layer deposited using optimized electropolymeriza- catalytic Properties of a Tetraruthenated Nickel Porphyrin. J. Chem.
tion procedure. Bioelectrochemistry 2007, 71, 46−53. Res. 1994, 1994, 1501−1515.
(24) Malinski, T.; Ciszewski, A.; Fish, J. R. Conductive Polymeric (42) Augusto, O. In CRC Handbook of Biomedicine of Free Radicals
Tetrakis(3-methoxy-4-hydroxyphenyl)porphyrin Film Electrode for and Antioxidants; Miguel, J., Ed.; CRC Press: Boca Raton, FL, 1989;
Trace Determination of Nickel. Anal. Chem. 1990, 62, 909−914. Vol. 3.
(25) Bailey, F.; Malinski, T. Carbon-Fiber Ultramicroelectrodes (43) Hohenberg, P. Inhomogeneous Electron Gas. Phys. Rev. 1964,
Modified with Conductive Polymeric Tetrakis(3-methoxy-4- 136, B864−B871.
hydroxyphenyl)porphyrin for Determination of Nickel in Single (44) Kohn, W.; Sham, L. J. Self-Consistent Equations Including
Biological Cells. Anal. Chem. 1991, 63, 395−398. Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133−
(26) Malinski, T.; Taha, Z. Nitric Oxide release from a Single Cell A1138.
Measured in Situ by a Porphyrinic-Based Microsensor. Nature 1992, (45) Becke, A. D. Density-functional thermochemistry. V. Systematic
358, 676−678. optimization of exchange-correlation functionals. J. Chem. Phys. 1997,
(27) Chen, S.-M.; Wang, C.-H. The electrocatalytic reactions of 107, 8554.
adenine, guanine, H2O, H2O2, N2H4, and L-cysteine catalyzed by (46) Becke, A. D. Density-functional thermochemistry. III. The role
poly(Ni(4-TMPyP)) film-modified electrodes. J. Solid State Electro- of exact exchange. J. Chem. Phys. 1993, 98, 5648.
chem. 2007, 11, 281−291. (47) Hill, E. A Problem in the Quantum Mechanics of Crystals. Phys.
(28) Porras Gutierrez, A.; Griveau, S.; Richard, C.; Pailleret, A.; Rev. 1931, 37, 785−794.
Gutierrez Granados, S.; Bedioui, F. Hybrid Materials from Carbon (48) Gouterman, M. Spectra of Porphyrins. J. Mol. Spectrosc. 1961, 6,
Nanotubes, Nickel Tetrasulfonated Phthalocyanine and Thin Polymer 138−163.
Layers for the Selective Electrochemical Activation of Nitric Oxide in (49) Araki, K.; Wagner, M. J.; Wrighton, M. S. Layer-by-layer growth
Solution. Electroanalysis 2009, 21, 2303−2310. of electrostatically assembled multilayer porphyrin films. Langmuir
(29) Danczuk, M.; Nunes, C. V., Jr.; Araki, K.; Anaissi, F. J. Influence 1996, 12, 5393−5398.
of Alkaline Cation on the Electrochemical Behavior of Stabilized Alpha (50) Buettner, G. R. Spin Trapping of the Hydroxyl Radical. In CRC
Nickel Hydroxide. J. Solid State Electrochem. 2014, 18, 2279−2287. Handbook of Methods for Oxygen Radical Research; Greenwald, R. A.,
(30) El-Shafei, A. A. Electrocatalytic oxidation of methanol at a nickel Ed.; CRC Press: Boca Raton, FL, 1985; pp 151−156.
hydroxide/glassy carbon modified electrode in alkaline medium. J. (51) Buettner, G. R. Spin trapping: ESR parameters of spin adducts.
Electroanal. Chem. 1999, 471, 89−95. Free Radical Biol. Med. 1987, 3, 259−303.
(31) Rocha, M. A.; Winnischofer, H.; Araki, K.; Anaissi, F. J.; Toma, (52) Buettner, G. R.; Mason, R. P. Spin-Trapping Methods for
H. E. A New Insight on the Preparation of Stabilized Alpha-Nickel Detecting Superoxide and Hydroxyl Free Radicals In Vitro and In
Hydroxide Nanoparticles. J. Nanosci. Nanotechnol. 2011, 11, 3985− Vivo. In Critical Reviews of Oxidative Stress and Aging: Advances in Basic
3996. Science, Diagnostics and Intervention; Rodriguez, R. G. C. a. H., Ed.;
(32) Martins, P. R.; Araújo Parussulo, A. L.; Toma, S. H.; Rocha, M. World Scientific: Singapore, 2003; Vol. I, pp 27−38.
A.; Toma, H. E.; Araki, K. Highly stabilized alpha-NiCo(OH)2 (53) Mottley, C.; Connor, H. D.; Mason, R. P. [17O] Oxygen
nanomaterials for high performance device application. J. Power Hyperfine Structure for the Hydroxyl and Superoxide Radials Adducts
Sources 2012, 218, 1−4. of Spin Traps DMPO, PBN AND 4-POBN. Biochem. Biophys. Res.
(33) Bukowska, J.; Rostonek, G.; Taraszewska, J. In-situ surface- Commun. 1986, 141, 622−628.
enhanced Raman spectroscopic study of gold electrodes modified with (54) Creutz, C.; Sutin, N. Reaction of tris(bipyridine)ruthenium(III)
nickel tetraazamacrocyclic complexes. J. Electroanal. Chem. 1996, 403, with hydroxide and its application in a solar energy storage system.
47−52. Proc. Natl. Acad. Sci. U.S.A. 1975, 72, 2858−2862.
(34) Fleischmann, M.; Hendra, P. J.; McQuillan, A. J. Raman spectra (55) Das, S. K.; Dutta, P. K. Intrazeolitic Photoreactions of
of pyridine adsorbed at a silver electrode. Chem. Phys. Lett. 1974, 26, Ru(bpy)33+ with Methyl Viologen. Langmuir 1998, 14, 5121−5126.
163−166. (56) Ghosh, P. K.; Brunschwig, B. S.; Chou, M.; Creutz, C.; Sutin, N.
(35) Czernuszewicz, R. S.; Macor, K. A. A Low-Temperature Bulk Thermal and Light-Induced Reduction of Ru(bipy)3 Aqueous
Electrolysis Cell for in situ Resonance Raman Spectroelectrochemis- Solution. J. Am. Chem. Soc. 1984, 106, 4772−4183.
try: Observation of the FeIVO Stretching Frequency for Electro- (57) Ledney, M.; Dutta, P. K. Oxidation of Water to Dioxygen by
generated Ferryl Tetramesitylporphyrin. J. Raman Spectrosc. 1988, 19, Intrazeolitic Ru(bpy)3. J. Am. Chem. Soc. 1995, 117, 7687−7695.
553−557. (58) Quayle, W. H.; Lunsford, J. H. Tris(2,2′-bipyridine)ruthenium-
(36) Frank, O.; Dresselhaus, M. S.; Kalbac, M. Raman Spectroscopy (III) in Zeolite Y: Characterization and Reduction on Exposure to
and in Situ Raman Spectroelectrochemistry of Isotopically Engineered Water. Inorg. Chem. 1982, 21, 97−103.
Graphene Systems. Acc. Chem. Res. 2015, 48, 111−118. (59) Li, X.-Y.; Czernuszewicz, R. S.; Kincaid, J. R.; Su, O.; Spiro, T.
(37) Salvatierra, R. V.; Moura, L. G.; Oliveira, M. M.; Pimenta, M. A.; G. Consistent Porphyrin Force Field. 2. Nickel Octaethylporphyrin
Zarbin, A. J. G. Resonant Raman spectroscopy and spectroelec- Skeletal and Substituent Mode Assignments from 15N, Meso-d4, and
trochemistry characterization of carbon nanotubes/polyaniline thin Methylene-d16, Raman and Infrared Isotope Shifts. J. Phys. Chem.
film obtained through interfacial polymerization. J. Raman Spectrosc. 1990, 94, 47−61.
2012, 43, 1094−1100. (60) Aydin, M. DFT and Raman spectroscopy of porphyrin
(38) Angnes, L.; Azevedo, C. M. N.; Araki, K.; Toma, H. E. derivatives: Tetraphenylporphine (TPP). Vib. Spectrosc. 2013, 68,
Electrochemical detection of NADH and dopamine in flow analysis 141−152.

4359 DOI: 10.1021/acs.langmuir.5b00250


Langmuir 2015, 31, 4351−4360
Langmuir Article

(61) Jiang, J.; Bao, M.; Rintoul, L.; Arnold, D. P. Vibrational


spectroscopy of phthalocyanine and naphthalocyanine in sandwich-
type (na)phthalocyaninato and porphyrinato rare earth complexes.
Coord. Chem. Rev. 2006, 250, 424−448.
(62) Li, X.-Y.; Czernuszewicz, R. S.; Kincaid, J. R.; Stein, P.; Spiro, T.
G. Consistent Porphyrin Force Field. 1. Normal-Mode Analysis for
Nickel Porphine and Nickel Tetraphenyiporphine from Resonance
Raman and Infrared Spectra and Isotope Shifts. J. Phys. Chem. 1990,
94, 31−47.
(63) Kieber-Emmons, M. T.; Annaraj, J.; Seo, M. S.; Heuvelen, K. M.
V.; Tosha, T.; Kitagawa, T.; Brunold, T. C.; Nam, W.; Riordan, C. G.
Identification of an “End-on” Nickel-Superoxo Adduct, [Ni(tmc)-
(O2)]+. J. Am. Chem. Soc. 2006, 128, 14230−14231.
(64) Reed, R. A.; Rodgers, K. R.; Kushmeider, K.; Spiro, T. G.; Su, Y.
O. Iron-Hydroxide Stretching Resonance Raman Bands of a Water-
Soluble Sterically Hindered Porphyrin. Inorg. Chem. 1990, 29, 2881−
2883.
(65) Yeo, B. S.; Bell, A. T. In Situ Raman Study of Nickel Oxide and
Gold-Supported Nickel Oxide Catalysts for the Electrochemical
Evolution of Oxygen. J. Phys. Chem. C 2012, 116, 8394−8400.

4360 DOI: 10.1021/acs.langmuir.5b00250


Langmuir 2015, 31, 4351−4360

You might also like