Sie sind auf Seite 1von 7

Journal of Inorganic Biochemistry 89 (2002) 54–60

www.elsevier.com / locate / jinorgbio

Redox behavior of melanins: direct electrochemistry of


dihydroxyindole-melanin and its Cu and Zn adducts
Shirley Gidanian, Patrick J. Farmer*
Department of Chemistry, University of California, Irvine, Irvine, CA 92697 -2025, USA
Received 30 July 2001; received in revised form 17 October 2001; accepted 5 November 2001

Abstract

Synthetic melanin films, formed on electrode surfaces by oxidative polymerization of 5,6-dihydroxyindole solution, were used to
directly measure the chromophore’s redox reactivity. Films on optically transparent indium-tin oxide (ITO) electrodes allow correlation of
spectral changes with electrochemical potential. Spectroelectrochemical titrations show an initial reversible transformation that is ascribed
to formation of a unique quinone-imine chromophore. The apparent E1 / 2 for maximum quinone-imine formation is |125 mV (vs.
Ag /AgCl) but at potentials higher than 100 mV, an irreversible bleaching is evident. Correlation of the current with the monomer
concentration implies that only one in six monomers is oxidized to the quinone-imine before the irreversible bleaching occurs. Films
pretreated with CuCl 2 and Zn(CH 3 COO) 2 show elevated quinone-imine absorbances, even under reducing conditions, indicating a
preferential stabilization of this state by coordination to the metals.  2001 Elsevier Science Inc. All rights reserved.

Keywords: Melanin; 5,6-Dihydroxyindole (DHI); Quinone-imine; Spectroelectrochemistry; Redox titration; Cu; Zn

1. Introduction within their cell walls [13]. Fe absorbance by melanin in


the brain is thought to promote Fenton-like formation of
Melanins are one of the largest classes of natural hydroxy radicals, and induce Parkinsonism [14]. Likewise,
pigments in the animal kingdom, commonly formed by melanin’s reaction with oxygen and other reactive oxygen
oxidative polymerization of catecholic species derived species are likely important in understanding their physio-
from tyrosine [1–3]. In humans, they are found in the hair logical roles [15,16].
and epidermis, as well as in areas of high nerve activity The chemical basis for melanin reactivity is its cate-
such as the middle ear, retina, and substantia nigra of the cholic subunits, dihydroxyindoles, which can undergo two
brain. They also may occur in a number of other tissues electron oxidations to the corresponding quinone form
either as a result of inflammation or as an important (Scheme 1). Melanins are subject to irreversible degra-
product of malignant tumors (usually termed melanomas) dation with a loss, or bleaching, of the characteristic
that arise from the melanin producing cells, the pigment color [16]. Unusual features of both native and
melanocytes [4]. An assumed physiological function of synthetic melanins are persistent radicals, as characterized
melanin is to buffer against photochemical stress, as the by an electron paramagnetic resonance (EPR) signal,
pigment is generated in response to UV damage [5]. But indicative of a sizable population of the semiquinone
melanins are also known to mediate oxidative stress by within the polymer [9,17,18].
neutralizing reactive oxygen species (ROS), such as As native melanins are in the form of colloidal particles,
superoxide and peroxide [6–8], and by absorbing metal they are difficult to examine using standard homogeneous
ions and radicals which promote oxidative damage [9–12]. solution techniques, and few details are known about the
Melanin’s redox activity has been linked to both pro- physicochemical nature of polymeric melanin’s redox
and anti-oxidant behavior, with important medical implica- activity [19]. Numerous electrochemical studies on
tions. For example, certain ‘black bacteria’ are protected eumelanin precursors and related catecholamines have
against phagocytic mechanisms by a layer of melanin been performed in which a darkening of the sample
solution around the anode has been proposed to indicate
*Corresponding author. Fax: 11-949-824-2210. melanin formation, but the colloidal particles thus formed
E-mail address: pfarmer@uci.edu (P.J. Farmer). are typically electrochemically silent [20]. Polymeric

0162-0134 / 01 / $ – see front matter  2001 Elsevier Science Inc. All rights reserved.
PII: S0162-0134( 01 )00405-6
S. Gidanian, P. J. Farmer / Journal of Inorganic Biochemistry 89 (2002) 54 – 60 55

Scheme 1. Speciation of monomers within DHI melanin.

melanin films have been shown to display unique electro- (HP E3617A) and the frequency data were recorded by a
chemical and semiconductor properties [21–23]. quartz crystal microbalance (QCM) measurement system
In this report, we follow the redox reactivity of polymer- (HP 8751A 5 Hz–500 MHz network analyzer) with
ized 5,6-dihydroxyindole (DHI) first reported by Horak LabVIEW 6.0 software.
and Weeks [24]. As made, the polymerized DHI (poly-
DHI) films have a bluish gray color, with a substantial 2.2. Poly-DHI film
absorbance throughout the visible range. Changes in the
absorbance with applied potential are used to follow All steps were performed in a glovebox under a nitrogen
reactivity within the melanoid film. By this method, a atmosphere. A DHI solution was prepared by hydrolyzing
direct measure of reversible and irreversible oxidations of DAI in situ: DAI (0.030 mmol, 6.0 mg) was dissolved in
the chromophore is obtained, as well as the effect of 12.5 ml of degassed ethanol to which 2 eq. of 0.1 M NaOH
absorbed Cu 21 and Zn 21 within the polymer. Our results (0.55 ml) needed to hydrolyze the acetoxy groups was
suggest reactivities assignable to specific components of added. This solution was allowed to stand for 1 h and then
that structure, as well as to the in-vivo properties of diluted up to 25 ml with pH 7.0, 0.05 M phosphate buffer.
various forms of melanin. Polymerization of DHI on electrode surface was achieved
using cyclic voltammetric deposition for 15 cycles at a
sweep rate of 50 mV s 21 for potentials varying from 20.40
2. Materials and methods to 0.35 V versus Ag /AgCl. Similar procedures were used
for deposition on gold or ITO substrates.
2.1. General experimental details
2.3. Spectroelectrochemical measurements
5,6-Diacetoxyindole (DAI) was obtained from TCI
America, Portland, OR. Indium-tin oxide plates were In a typical procedure, a poly-DHI film deposited on
purchased from Quantum Coating, Mt Laurel, NJ. All ITO plate was placed in a three-neck, anaerobic UV-visible
other chemicals were reagent grade quality. Mass spectra cell containing 0.1 M LiCl / 0.05 M phosphate buffer pH
were obtained on a VG analytical 707E instrument. Laser 7.0 solution as electrolyte. After obtaining the absorbance
desorption ionization (LDI) measurements were performed spectrum of the film without applied potential as a
on a Perspective Biosystem DE STR time-of-flight instru- reference, the ITO electrode was connected to the electro-
ment, operating in positive-ion linear mode. Ions were chemical analyzer as working electrode, using a platinum
formed by a pulsed UV laser beam (nitrogen laser, l5337 wire as counter electrode and an Ag /AgCl reference
nm). Absorbance spectra were obtained on a Hewlett- electrode. Variable potentials were applied to the film and
Packard 8453 UV/ visible spectrometer. Voltammograms the absorbance spectrum of the film was obtained. Subtrac-
were obtained with commercial gold or hand cut indium- tion of the reference spectrum from these spectra provided
tin oxide (ITO) plates. A three-electrode voltammetric cell the background subtracted spectra.
was used with a platinum counter electrode and Ag /AgCl
as reference electrode. Cyclic voltammetry experiments 2.4. Current efficiency and stoichiometry
were carried out with a BAS-100B-W electrochemical
analyzer. Anaerobic experiments were carried out in a In order to quantitate the electron efficiency of the
glovebox, or by purging the analyte solution with nitrogen polymerization process, films were made by bulk electro-
for at least 30 min. All reported potentials are versus chemical oxidation of DHI solution onto a gold plated
Ag /AgCl reference electrode. quartz crystal microbalance, and the mass-induced fre-
AT-cut quartz crystals with a basic resonant frequency quency shift determined. In a typical experiment, poly-DHI
of 5 MHz were purchased from Colorado Crystal Corpora- film was deposited on one side of the quartz electrode by
tion. The quartz plates (15 mm diameter) with gold bulk electrolysis of the DHI solution at 20 mV, after
electrodes (8.5 mm diameter) were fixed in the measuring recording the original frequency of the crystal. The total
cell made of aluminum whose temperature was controlled current passed through the electrode was 4.415310 23 C,
by a heater (DALE RH-10) attached to a DC power supply which corresponds to 2.287310 28 mol of DHI deposited.
56 S. Gidanian, P. J. Farmer / Journal of Inorganic Biochemistry 89 (2002) 54 – 60

The crystal was then rinsed with water and ethanol, dried
under nitrogen, and its frequency was measured. A 60-Hz
frequency shift was detected which correspond to 1.2 mg
(0.805 mol) of the material deposited on the surface.
Further deposition experiments were done on both gold
and ITO electrode surface at 20 mV versus Ag /AgCl. The Scheme 2. Oxidative deposition of DHI film.
resulting films were then rinsed several times with deion-
ized water and stored in the glovebox. mechanically stable and insoluble in water or other organic
Bulk electrolysis of the deposited films allowed quantifi- solvents tested.
cation of the oxidative current passed at various potentials The voltammetric response of the poly-DHI films is
(after correction for the deposition efficiency). In a typical slow due to limited charge and ion transport through the
set of experiments, a deposition current of 2.7310 23 C multi-layer film; only broad, low-current waves are seen
was passed to form a film on Au; whereas, after cleaning, for the polymer itself at normal scan rates above 50 mV
only 19.7310 25 C were passed during stepwise bulk s 21 . Cyclic voltammograms of the film at very slow rates
electrolysis at 50-mV intervals from 2100 to 50 mV, and (,1 mV s 21 ) show an oxidation at 100 mV versus Ag /
1.1310 23 C total from 2100 to 300 mV. These data were AgCl (Fig. 2). The reduction peak is broad and small,
used to calculate the current passed during the spectrally- indicating a depletion of the redox active component
characterized reversible oxidation, and the subsequent following oxidation.
irreversible oxidations.
3.2. Spectroelectrochemical measurements
2.5. Mass spectral characterization
Though highly colored, thin films of DHI-melanin are
To obtain mass spectral characterization, poly-DHI films transparent, and characterizable by their absorption spectra.
were electrodeposited directly onto stainless steel targets Spectroelectrochemical redox titrations of films deposited
suitable for use with the MALDI–TOF instrumentation on optically transparent film indium-tin oxide (ITO)
using cyclic voltammetry technique, by connecting the electrodes were performed to characterize the change in
sample holder to electrochemical analyzer as the working absorbance with applied potential.
electrode. Fig. 3 illustrates the spectral change of a poly-DHI film
as the applied potential is varied from 250 mV versus
2.6. Cu- and Zn-treated poly-DHI Ag /AgCl to 150 mV. An increase in absorbance is seen at
500 nm, and a decrease at 380 nm, with an isosbestic point
Poly-DHI films made by cyclic voltammetry on ITO at 417 nm which implies a clean interconversion of
electrodes were soaked in a 0.1 M CuCl 2 , or 0.1 M species. Under anaerobic conditions, this absorbance
Zn(CH 3 COO) 2 solution under nitrogen atmosphere over- change is reversible and can be repeated several times. The
night. The electrodes were then rinsed with deionized difference spectra in Fig. 3 are similar, in both maxima and
water. Spectroelectrochemical experiments on the Cu- or
Zn-treated ITO electrodes were as stated above.

3. Results and discussion

3.1. DHI-melanin films

To measure the redox behavior of colloidal melanins,


we mimic biological melanogenesis by oxidative poly-
merizing melanin onto electrodes. Voltammetric scans of
an aqueous solution of DHI initiates the polymerization
and deposits a film of poly-DHI on the working electrode
surface (Scheme 2) [24].
The sequential voltammograms obtained during deposi-
tion are shown in Fig. 1. The anodic peak of the first scan
corresponds to the oxidation of DHI on a clean electrode
surface; subsequent cycles display the blocking of the DHI
oxidation and a reductive component due to the growing Fig. 1. Cyclic voltammetric polymerization of 5,6-dihydroxyindole (1.0
poly-DHI film. Film thickness can be controlled by the mM in pH 7.0, 50 mM phosphate buffer) at scan rate of 50 mV s 21 on
scan rate and number of cycles. Once formed, the films are small gold disk electrode for ten cycles.
S. Gidanian, P. J. Farmer / Journal of Inorganic Biochemistry 89 (2002) 54 – 60 57

Fig. 2. Cyclic voltammogram of a poly-DHI film in 0.1 M KCl and pH


7.0, 0.1 M phosphate buffer at 0.5 mV s 21 scan rate.
Fig. 4. Open-circuit subtracted spectra of poly-DHI film on ITO demon-
strating irreversible bleaching. Sequential spectra obtained after applying
potentials of: (a) 2300 mV; (b) 2100 mV; (c) 0 mV; (d) 100 mV; (e) 300
mV; (f) 500 mV; and (g) 600 mV. Conditions as described in the text.

lost, indicating a change in the chemical composition (Fig.


4). The irreversible changes are more dramatic at po-
tentials above 1300 mV, where irreversible bleaching is
much increased at higher wavelength, e.g. 700 nm (Fig. 4).
Plotting the absorbance change versus applied potential
gives the apparent E1 / 2 potential for the quinone-imine
formation at |125 mV versus Ag /AgCl (Fig. 5). The loss
of reversibility above 50 mV suggests that the chemical
decomposition is dependent on the concentration of quin-
one-imine monomers.
The absorbance of an as-made film typically shows
substantial quinone-imine absorbance, and can be used to
estimate the resting redox state of the melanin. These
Fig. 3. Open-circuit subtracted absorbance spectra of poly-DHI film on resting potentials vary somewhat from sample to sample,
ITO demonstrating reversibility of oxidation. Sequential spectra obtained but are generally in the range of 0 to 50 mV, i.e. reversibly
after applying potentials of: (a) 250 mV; (b) 0 mV; (c) 250 mV; and (d) 0
oxidized. The film’s resting redox state is also affected by
mV. Conditions as described in the text.
its age and history, e.g. exposure of the dry films to oxygen

isosbestic point, to those obtained during chemical oxida-


tion of L-dopa, attributed to the formation of a quinone-
imine intermediate, dopachrome [26,27]. Thus the revers-
ible spectral changes are due to formation of such quinone-
imine functionalities within the DHI-melanin, formed by
tautomerization from the two-electron oxidized quinone
state (Scheme 3).
At potentials in the range from 150 to 1300 mV, this
characteristic absorbance increases, but the reversibility is

Fig. 5. Plot of absorbance versus applied potential for poly-DHI film on


Scheme 3. Tautomerization of quinone to quinone imine. ITO at (a) 500 nm and (b) 700 nm.
58 S. Gidanian, P. J. Farmer / Journal of Inorganic Biochemistry 89 (2002) 54 – 60

slowly increases their observed resting potential. For this


reason, films were made and stored under an inert atmos-
phere. The presence of oxygen has a dramatic effect on the
absorbance changes during spectroelectrochemical elec-
trolysis, and will be the subject of a subsequent report.

3.3. Quantification of the reversible and irreversible


oxidations

To determine the number of electrons per monomer


during chromophore formation required a measurement of
both the current and the number of monomers. The current
efficiency for deposition was determined using a quartz
crystal microbalance to measure the mass of DHI-melanin
deposited on the electrode surface. The expected oxidative
current needed for high poly-DHI polymer formation Fig. 6. Plot of absorbance versus applied potential for (a) poly-DHI film,
(b) Cu-treated poly-DHI film, and (c) Zn-treated poly-DHI film at 500
corresponds to two electrons per monomer. The deposition
nm.
current was kept small, to be comparable to the thin films
used in spectroelectrochemical measurements, and was
therefore likely to contain a high component of non- The total oxidative charge during the quinone-imine
Faradaic background current. Under these conditions, the chromophore formation (between 250 and 1300 mV)
current efficiency for deposition over a set of experiments was determined to be |1.1 electrons per monomers, about
was calculated to be 44%, corresponding to a surface half that required for a stoichiometric conversion of
coverage of |8.75 nmol / cm 2 for a close-packed film of
quinole to the quinone. Thus the buffering capacity of
DHI monomers (lower limit |18 layers of poly-DHI).
melanins in this range is much less than expected.
Using these films, the current observed during the revers-
ible transformation corresponded to only 16% of the total
3.4. Effect of absorbed Cu and Zn
expected oxidation current. By this measure, only a small
number of monomers are reversibly oxidized before ir-
It is known that melanins absorb metal ions in vivo [30]
reversible changes occur.
and these metals may dramatically alter the redox reactivi-
Laser desorption / ionization mass spectrometry (LDI-
MS) of the poly-DHI film displays repeating peak clusters, ty of the polymer [16]. The effect of absorbed metal ions
suggesting the sample is composed of oligomers rather on the melanin films’ spectroelectrochemical behavior was
than long polymeric chains. Peak clusters with m /z differ- assayed by pretreatment with metal ion solutions. The
ences of 147 mass units are typical, and can be ascribed to redox titration results for poly-DHI sample as well as
cationic species originating from trimers up to dodecamers Zn 21 - and Cu 21 -treated samples are shown in Fig. 6,
of DHI. Similar clusters have been seen for colloidal obtained by first fully reducing the films and varying the
DHI-melanin solutions, and for a number of synthetic and potential positively.
native melanins [25,28,29]. Quantification of the polydis- Metal-treated samples show a dramatic increase in
persity of the oligomers is difficult, as desorption selective- absorbance at 500 nm, which varies by metal, and is
ly favors the smaller chained oligomers. Microscopy evident before applying potential. This is likely due to the
studies have suggested a fundamental repeating unit of effect of the metal ions on the equilibrium between
synthetic melanins, melanin protomolecules, modeled as quinone and quinone-imine functionalities in the film
randomly-coupled pentamers and hexamers of DHI (Scheme 4). The absorbance of Zn-treated poly-DHI film
subunits, as illustrated below. Using this model, the at open circuit corresponds to an applied potential of |200
reversible oxidation corresponds to about two electrons mV for metal-free films, and for Cu-treated samples the
(one quinone-imine) per protomolecule. absorbance is somewhat above that obtainable before the
irreversible bleaching occurs, implying a high concen-
tration of quinone-imine. Reduction induces the loss of
weakly bound ions, as evident by a loss of absorbance. But

Scheme 4. Effect of metal binding on quinone tautomerization.


S. Gidanian, P. J. Farmer / Journal of Inorganic Biochemistry 89 (2002) 54 – 60 59

after the initial reduction, the reversibility of electrochemi- The formation of the quinone-imine tautomers is pro-
cal oxidation of both Cu- and Zn-treated films are similar moted by absorption of metal ions such as Zn(II) and
to that of the metal-free samples, indicating a stable Cu(II). The metal-bound quinone-imines are then inert to
concentration of bound metal. The maximum concentration reduction, but not to the irreversible bleaching upon
of quinone-imine obtainable remains relatively constant for oxidation. The effect of the metal-ion binding on the
all samples, suggesting that the decomposition pathway is oxidation potential of the melanin will depend on the
dependent on the quinone-imine concentrations and in- binding properties of the metal, and may substantially vary
sensitive to metal content. the potential at which the oxidation occurs.
Complete reduction (to 2500 mV) only partially de-
creases the absorbance of metal-treated DHI-melanins. The
Zn-treated samples retain 50% of the highest quinone- 5. Abbreviations
imine absorbance, the Cu-treated sample close to 70%,
roughly in line with the affinity of the metal ions for DAI 5,6-diacetoxylindole
melanin [31]. This indicates that binding of metals makes a DHI 5,6-dihydroxyindole
population of quinone-imine functionalities inert to electro- EPR electron paramagnetic resonance
chemical reduction, and therefore limits the redox buffer- ITO indium-tin oxide
ing ability of the melanin. LDI laser desorption ionization
Metal-binding may also effect the oxidation potential of QCM quartz crystal microbalance
the melanins. As derived from redox titrations in Fig. 6, ROS reactive oxygen species
the apparent E1 / 2 for poly-DHI and for the Cu-treated
sample are close (150 and 125 mV), but that for the
Zn-treated sample is much lower, |0 mV. Repeated
experiments show that Zn-treatment makes the melanin Acknowledgements
substantially easier to oxidize. The shift of potential
implies a preference for Zn-coordination to the oxidized We thank Dr Frank Meyskens for his interest and
quinone-imine form, which would perturb the redox inspiration, Dr M. Bayachou and Michelle Tran for their
equilibrium towards that state under these conditions [32]. initial efforts on this project, and Dr Bruno Szpoganicz for
The small shift seen in Cu-treated samples would likewise thoughtful discussions and interpretations. Dr Peter
suggest an equal affinity of both the quinole and quinone Taborek is thanked for the QCM analysis. This work was
states for Cu-binding in the electroactive DHI monomers. supported by grants from the Chao Family Cancer Center.

4. Conclusions References

Depositing a melanin film on an electrode surface offers [1] G. Prota, in: G. Prota (Ed.), Melanins and Melanogenesis, Academic
several advantages in studying its chemical reactivity: (i) Press, San Diego, 1992, pp. 153–184.
the films are easy to prepare and handle as compared to [2] R. Crippa, V. Horak, G. Prota, P. Svoronos, L. Wolfram, in: A. Brossi
colloidal melanin solutions; (ii) they can be formed on (Ed.), Alkaloids, Academic Press, New York, 1989, pp. 253–323.
[3] P.A. Riley, Int. J. Biochem. Cell Biol. 29 (19) (1997) 1235–1239.
transparent supports which allows coupled electrochemical
[4] G. Prota, M. d’Ischia, D. Mascagnena, Melanoma Res. 4 (1994)
and spectroscopic analysis; and (iii) they allow a platform 351–358.
for following reactions with solution-based species, such [5] N. Kollias, R.M. Sayre, L. Zeise, R. Chedekel, J. Phototchem.
as metal absorption by the melanin. Photobiol. B Biol. 9 (1991) 135–160.
Using this simple method we have obtained direct [6] F.L. Meyskens, P.J. Farmer, J. Fruehauf, Pigment Cell Res. 14
(2001) 148–154.
measurements of DHI-melanin’s oxidation, and examined
[7] W. Koritowski, P. Hintz, R.C. Sealy, B. Kalayanaraman, Biochem.
the effect of absorbed metal ion on this behavior. The Biophys. Res. Commun. 131 (2) (1985) 659–665.
oxidation of melanin proceeds in two distinct steps, the [8] M. Rozanowska, T. Sarna, E.J. Land, T.G. Truscott, Free Radic.
first of which is reversible, but which corresponds to Biol. Med. 26 (1999) 518–525.
formation of one quinone per six monomers. The spectral [9] R.C. Sealy, C.C. Felix, J.S. Hyde, H.M. Swarz, W.A. Pryor, in: Free
Radicals in Biology, Vol. IV, Academic Press, New York, 1980, pp.
changes associated with the reversible oxidation are inter-
210–260.
preted as formation of quinone-imine functionalities. The [10] B. Larsson, H. Tjalve, Acta Physiol. Scand. 104 (1978) 479–484.
second reactivity is an irreversible bleaching reaction, [11] T. Sarna, W. Froncisz, J.-S. Hyde, Arch. Biochem. Biophys. 202 (1)
beginning after substoichiometric oxidation, with a maxi- (1980) 289–303.
mum of |50% conversion of quinole to the quinone-imine. [12] L. Novellino, A. Napolitano, G. Prota, Chem. Res. Toxicol. 12
(1999) 985–992.
Thus natural melanins likely act as disposable buffers of
[13] E.S. Jacobson, N.D. Jenkins, J.M. Todd, Infect. Immunol. 62 (1994)
oxidative stress, in that localized decomposition of mono- 4085–4086.
mers occurs upon oxidation, rather than wholescale con- [14] D. Ben-Shachar, M.B.H. Youdim, Prog. Neuropsychopharmacol.
version of quinoles to quinones. Biol. Psychiatry 17 (1993) 139–150.
60 S. Gidanian, P. J. Farmer / Journal of Inorganic Biochemistry 89 (2002) 54 – 60

[15] T. Sarna, H.M. Swarz, in: G. Scott (Ed.), Atmospheric Oxidation [24] V. Horak, G. Weeks, Bioorg. Chem. 21 (1993) 24–33.
and Antioxidants, Vol. III, Elsevier, Amsterdam, 1993, pp. 129–165. [25] A. Napolitano, A. Pezzella, G. Prota, R. Seraglia, P. Traldi, Rapid
[16] W. Kortawski, T. Sarna, J. Biol. Chem. 265 (21) (1990) 12410– Commun. Mass Spectrom. 10 (1996) 468–472.
12416. [26] J.N. Rodrigues-Lopez, J. Tudela, R. Varon, F. Garcia-Canovas,
[17] T. Sarna, J.S. Hyde, H.M. Swartz, Science 192 (1976) 1132–1134. Biochim. Biophys. Acta 1076 (1991) 379–386.
[18] R.C. Sealy, J.S. Hyde, C.C. Felix, I.A. Menon, G. Prota, Science [27] J. Cabanes, F. Garcia-Canovas, J.A. Lazano, F. Garcia-Carmona,
217 (1982) 545–547. Biochim. Biophys. Acta 923 (1987) 187–195.
[19] T. Sarna, H.M. Swartz, in: J.J. Nordlund, R.E. Boissy, V.J. Hearing, [28] A. Bertazzo, C.V.L. Costa, G. Allegri, D. Favretto, P. Traldi, J. Mass
R.A. King, J.P. Ortone (Eds.), The Pigmentary System: Physiology Spectrom. 34 (1999) 922–929.
and Pathophysiology, Oxford University Press, New York, 1998, pp. [29] A. Pezzella, M. d’Ischia, A. Napolitano, A. Palumbo, G. Prota,
333–358. Tetrahedron 33 (24) (1997) 8281–8286.
[20] G. Dryhurst, Chem. Rev. 90 (1990) 795–811. [30] C. Sarzanini, E. Mentasti, O. Abollino, M. Fasano, Mar. Chem. 39
[21] G.M. Robinson, E.I. Iwuoha, M.R. Smyth, Electrochim. Acta 34 (1992) 243–250.
(1998) 3489–3496. [31] B. Szpoganicz, S. Gidanian, P. Kong, P.J. Farmer, J. Inorg. Biochem.
[22] J. McGinness, P. Corry, P. Proctor, Science 183 (1974) 853–855. 89 (2002) 45–53.
[23] M.M. Jastrzebska, S. Jussila, H. Isotalo, J. Mater. Sci. 33 (1998) [32] R.C. Kapoor, B.S. Aggarwal, in: Principles of Polarography, Wiley,
4023–4028. New York, 1991, pp. 54–65, Chapter 5.

Das könnte Ihnen auch gefallen