You are on page 1of 8

Electrochimica Acta 81 (2012) 197–204

Contents lists available at SciVerse ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Electrochemical and chemical formation of a low-barrier proton transfer


complex between the quinone dianion and hydroquinone
Pablo D. Astudillo a , Drochss P. Valencia a , Miguel A. González-Fuentes a , Blanca R. Díaz-Sánchez a ,
Carlos Frontana b , Felipe J. González a,∗
a
Departamento de Química del Centro de Investigación y de Estudios Avanzados del I.P.N., Apartado Postal 14-740, 07360, México, D.F., Mexico
b
Centro de Investigación y Desarrollo Tecnológico en Electroquímica, Parque Tecnológico Querétaro, Sanfandila, 76703, Pedro Escobedo, Querétaro, Mexico

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

Article history: The electrochemistry of the 1,4-benzoquinone and hydroquinone in acetonitrile and dimethylsulfox-
Received 19 April 2012 ide was reviewed to explain the nature of a broad reversible signal that appears during the reduction
Received in revised form 21 July 2012 of quinones with small amounts of proton donors and that cannot be explained in the framework of
Accepted 21 July 2012
the classical mechanisms of quinone reduction. Cyclic voltammetry and NMR experiments as well as
Available online 27 July 2012
electronic structure calculations were performed to show that under specific conditions, the anion QH−
disproportionates into a face-to-face dianionic quinhydrone associated by strong intermolecular hydro-
Keywords:
gen bonds. This complex explains the mentioned broad signal and the results presented show that it can
Quinone
Quinhydrone
be formed during the reduction of benzoquinone in the presence of stoichiometric amounts of weak pro-
Proton transfer ton donors such as acetic acid and hydroquinone, or by half-deprotonation of hydroquinone. An analysis
Hydrogen bonding of the energetic barriers of the hydroxyl protons involved in the complex is also presented in this work.
Charge transfer complex © 2012 Elsevier Ltd. All rights reserved.

1. Introduction of electrochemical studies with the aim to understand the differ-


ent molecular factors determining the stability or reactivity of the
The quinone moiety is a redox active cofactor whose bidirec- reduced species. The studies on the electrochemical behaviour of
tional conversion into hydroquinone plays a key role in biological quinones have been performed in buffered [23] and unbuffered [24]
electron and proton transport processes, such as those occurring aqueous media as well as in polar organic solvents [25,26].
at mitochondria, chloroplasts and bacterial cell membranes [1–4]. It is known that the basicity and nucleophilic power of the
Important examples of the biological activity of these lipid soluble reduced quinonoid species are intrinsically related with the final
electron and proton carriers are found in literature, for instance, reactivity of these intermediates. These chemical properties can
in the redox cycle of vitamin K [5–7], in the respiratory chains be modified by the presence of electron-releasing or electron-
of aerobic organisms where mitochondrial NADH:ubiquinone oxi- withdrawing substituents on the quinone structure [27–29] or
doreductase (complex I), cytochrome bc1 oxidoreductase (complex through the establishment of intramolecular hydrogen bonds,
II) and succinate-quinone reductase (complex III) are important such as in the case of ␤-hydroxynaphthoquinones [30–34]. For
components [8–11] and also in the process of photosynthesis in quinones not substituted with acidic functional groups (Q), the
chloroplasts where plastoquinone or phylloquinone are the redox electrochemical reduction in aprotic solvents is carried out fol-
cofactors [12–14]. The electron transfer properties of quinones lowing two quasireversible consecutive one-electron transfer steps
have also important implications in cellular defense since they [35]. Under these conditions, the electrochemically generated
show antibacterial [15,16], antifungal [17,18] and antiparasitic semiquinone radical anion (Q•− ) and the dianion (Q2− ) are sta-
properties [19,20]. It is also remarkable that the quinone structure ble and both can be directly detected on the reverse scan in
and the functional groups involved in the quinonoid structure are cyclic voltammetry experiments [25]. The structural properties
both essential to develop specific biochemical functions such as of these fundamental intermediates have been established from
cytotoxicity and anticancer activity [21,22]. electrochemical methods coupled with UV–vis [36,37], FTIR [38],
The importance of the quinone-to-hydroquinone bidirectional ESR [39,40] and NMR [40] spectroscopies. As an external fac-
conversion in biological systems has motivated a great number tor contributing to the stabilization of these intermediates, the
association with metallic cations [41–43] and weak hydrogen
bonding donors has been demonstrated for many quinonoid
∗ Corresponding author. Fax: +52 55 57473389. systems [36,44–47]. Additionally, stabilization effects by ␲–␲
E-mail address: fgonzale@cinvestav.mx (F.J. González). type charge transfer interactions have also been reported in

0013-4686/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.electacta.2012.07.078
198 P.D. Astudillo et al. / Electrochimica Acta 81 (2012) 197–204

the case of the dianion species and aromatic hydrocarbons


[48].
When the reduction process of quinones is carried out in apro-
tic solvents in presence of an excess of moderately strong proton
donor (DH) such as carboxylic acids, the quinone is directly trans-
formed into hydroquinone QH2 , which is strongly associated with
the conjugated base of the proton donor D− [49]. The generally
accepted mechanism for a wide variety of quinones and proton
donors involves two electron and two proton transfer reactions,
which occur in the framework of the classical square scheme (ECEC)
competing with first or second order disproportionation mecha-
nisms (DISP1 or DISP2) [25,50].
The transition from the one electron transfer process in apro-
tic medium towards the two-electron and two-proton transfer
mechanism in acidic medium has revealed new insights about the
overall mechanism of the quinone-to-hydroquinone bidirectional
conversion. For example, when the concentration of the proton
donor is close to the quinone concentration, a broad peak between
those corresponding to the formation of semiquinone and quinone
dianion has been observed [51]. The species involved in the clas-
sical electron and proton transfer scheme are not able to explain
the appearance of this broad peak. However, the intervention of Fig. 1. Cyclic voltammetry of 1,4-benzoquinone 2 mM in the presence of several
concentrations of acetic acid, in acetonitrile + n-Bu4 NPF6 0.2 M, on a glassy carbon
an anionic complex, analogue to the well known charge trans-
electrode ( = 3 mm) at 0.1 V s−1 .
fer system quinhydrone [52], was consistent with this unexpected
peak. Thus, in the present work, the nature of this anionic type-
quinhydrone complex, which is a good example of a low-barrier electrode; this electrode was polished with 1 ␮m alumina powder
proton transfer complex, was studied in acetonitrile and dimethyl- (Buehler) and ultrasonically rinsed with anhydrous ethanol before
sulfoxide. Three ways to prepare and analyse the proposed complex each run. The counter electrode was a platinum mesh and the ref-
were considered: (1) in situ formation during the reduction of 1,4- erence electrode an aqueous saturated calomel electrode (SCE). A
benzoquinone in the presence of sub-stoichiometric amounts of salt bridge, containing a solution of 0.1 M n-Bu4 NPF6 in either ace-
acetic acid, (2) formation during the association of hydroquinone tonitrile or DMSO, was used to connect the cell with the reference
with the quinone dianion generated by electrochemical reduction electrode.
1 H and 13 C NMR spectra were recorded on a Jeol GSX-270 spec-
of benzoquinone and, (3) by half-deprotonation of hydroquinone.
These studies were carried out by cyclic voltammetry, 1 H NMR trometer in deoxygenated deuterated acetonitrile.
spectroscopy and electronic structure calculations.
2.3. Calculations
2. Experimental
Geometry optimization and frequency calculations of the
dianion–hydroquinone and monoanionic–monoanionic complexes
2.1. Chemicals and procedures
were carried out with the program Gaussian 09 Revision B.01 [53],
using the approach of the Density Functional Theory. Four differ-
1,4-Benzoquinone 99%, hydroquinone 99%, acetic acid 99%,
ent functionals were employed: BHandHLYP [54], MPW1PW1 [55],
tetrabutylammonium hydroxide 1 M/methanol and tetraethy-
MPW1PBE [56] and the long range corrected PBE functional, LC-
lammonium hydroxide 1 M/water were Aldrich chemicals. 1,4-
PBE [57–60], with a 6-311++G(d,p) basis set. Frequency analysis
Benzoquinone was previously purified by sublimation under
for all the involved structures were performed after full geometry
reduced pressure. Acetonitrile and dimethylsulfoxide spectropho-
optimizations, revealing the absence of negative frequencies, thus
tometric grade (Merck Uvasol) were the solvents used for the
indicating that the structures are minimum energy conformers.
electrochemical and spectroscopic experiments and they were
used as received. Tetrabutylammonium hexafluorophosphate 98%
(Aldrich), recrystallized from absolute ethanol and vacuum-dried 3. Results and discussion
overnight in a Schlenck tube at ∼80 ◦ C, was used as the sup-
porting electrolyte. The hydroquinone deprotonation was carried 3.1. Electrochemical reduction of 1,4-benzoquinone in presence
out in inert atmosphere by using tetrabutylammonium hydroxide. of acetic acid
The chemical and electrochemical experiments were performed at
room temperature (∼25 ◦ C). All the working solutions were deoxy- The electrochemical behaviour of 1,4-benzoquinone (Q) in ace-
genated with dry argon bubbling directly to the bulk solution before tonitrile, in absence and in presence of acetic acid (HAc), is shown
every voltammetric scan, keeping the inert atmosphere in the cell in Fig. 1. In aprotic media (Fig. 1a), two typical consecutive one-
during each experiment. electron quasi-reversible peaks, affording the radical anion Q•−
and dianion Q2− , were observed. Peaks Ic–Ia (E1o = −0.498 V vs SCE)
correspond to the redox couple Q/Q•− (Eq. (1)) while peaks IIc–IIa
2.2. Instrumentation, cell and electrodes (E2o = −1.239 V vs SCE) correspond to the couple Q•− /Q2− (Eq. (2))
[25,35].
A DEA-332 potentiostat (Radiometer, Copenhagen) was used
for all the electrochemical experiments applying positive feedback Eo •−
1
Q + e− Q (1)
resistance compensation. Voltammetric experiments were carried
out on a conventional three-electrode cell. A 3 mm diameter glassy •−
Eo
− 2
carbon disk (Sigradur G, from HTW Germany), was used as working Q + e Q2− (2)
P.D. Astudillo et al. / Electrochimica Acta 81 (2012) 197–204 199

Alternatively, in the presence of an excess of acetic acid (Fig. 1d),


the second reversible reduction wave IIc–IIa disappears while the
first reversible wave Ic–Ia becomes a less negative chemically irre-
versible wave IIIc (Ep = −0.245 V vs SCE at 0.1 V s−1 ) that involves
the well known two-electron and two-proton transfer mechanism
[25], which occurs in the framework of the classical competition
between the ECEC (Eqs. (1), (3)–(5)) and disproportionation (Eqs.
(1), (3), (4 ), and (5)) mechanisms [61]. The disappearance of wave
IIc–IIa as well as the transition of wave Ic into wave IIIc is in agree-
ment with the fact that QH• is more easily reduced than Q (E4o > E1o ).
As a result of this condition, the current peak of wave IIIc corre-
sponds approximately to a two-fold increase compared with the
peak current of wave Ic.
•−
Q + e−  Q

•− •
Q + HAc  QH + Ac− (3)

QH + e−  QH− (4)

and/or
• •−
QH + Q  QH− + Q (4 )

QH− + HAc  QH2 + Ac− (5)

Q + 2e− + 2HAc = QH2 + 2Ac− (6) Fig. 2. Cyclic voltammetry of hydroquinone 2 mM in the presence of tetrabutylam-
monium hydroxide, in acetonitrile + 0.2 M n-Bu4 NPF6 , on glassy carbon electrode
(3 mm ␾) at 0.1 Vs−1 . The concentration of n-Bu4 NOH was (A) 0 mM, (B) 2 mM and
Due to the fact that acetate ion (Ac− ) and hydroquinone (QH2 )
(C) 4 mM.
are simultaneously produced in the overall reaction mechanism
of Eq. (6) (Eqs. (1) + (3) + (4) + (5) or (1) + (3) + (4 ) + (5)), a hydrogen
bonding association complex, QH2 (Ac− )2 , is formed. This complex 3.2. Chemical formation of a quinone dianion–hydroquinone
is oxidized at the level of peak IIIa (Fig. 1), whose oxidation peak complex
potential can range from −0.078 to 0.241 V vs SCE depending on
the acetic acid concentration. The magnitude of this potential is In order to understand the nature of peaks IVc–IVa, the electro-
lower than that of free hydroquinone under the same experimental chemical behaviour of pure hydroquinone QH2 in acetonitrile was
conditions (Ep = 1.027 V vs SCE at 0.1 V s−1 ) [49]. recorded (Fig. 2A). The typical two-electron oxidation peak IIIa [49]
During the transition from one-electron towards two-electron was observed, which is chemically irreversible and is located at a
transfer mechanisms, the most interesting feature of the voltam- peak potential value of Ep = 1.027 V vs SCE at 0.1 V s−1 . Signal IIIc is
metric behaviour was obtained at stoichiometric concentrations of related to reduction of protonated 1,4-benzoquinone, which is the
HAc ([HAc]/[Q] ≤ 1) (Fig. 1b and c). Under these conditions, the last product formed at the level of peak IIIa [49].
protonation step of the mechanism (Eq. (5)) cannot be favoured Fig. 2B shows the voltammogram corresponding to the product
and consequently the anion QH− should be the only expected of half-deprotonation of QH2 , which in principle, should correspond
species, such as depicted by the overall Eq. (7) (Eqs. (1) + (3) + (4) to the anion QH− . Starting from the highly negative open circuit
or (1) + (3) + (4 )). potential value (ocp = −0.938 V vs SCE) and scanning towards the
In the literature, an oxidation peak like IIIa has been sometimes positive direction, the anodic segment of the broad peak IVa was
assigned to the species QH− [26]. However, the problem with this found to be coincident with peak IVa of Fig. 1C, obtained by reduc-
proposal arises from the fact that this oxidation peak corresponds tion of Q in presence of stoichiometric amounts of HAc. As shown in
rather to the voltammetric response of the association complex Fig. 2B, the oxidation (at the level of peak IVa) of the product of half-
between QH2 and the conjugated base of acetic acid [49]. Therefore, deprotonation of hydroquinone (in principle a species derived from
an alternative process is occurring with QH− at stoichiometric con- QH− ) gives rise to semiquinone Q•− (peak Ia) and hydroquinone
centrations of HAc and this must be related to the presence of the QH2 (peak IIIa). This result and the fact that the peak potential of
unexpected broad peaks IVc–IVa (Epc = −0.831 and Epa = −0.724 V IVa (Ep = −0.724 V vs SCE at 0.1 V s−1 ) is quite negative and it occurs
vs SCE) in Fig. 1, which appears between the redox signals of Q/Q•− at a potential similar to that for the oxidation of quinone dianion
(Ic–Ia) and Q•− /Q2− (IIc–IIa). Q2− (peak IIa in Fig. 2C) (Epa = −0.881 V vs SCE at 0.1 V s−1 ), allows
At this point, the main feature of the results carried out at to propose that QH− is not stable under the experimental condi-
sub-stoichiometric and stoichiometric concentrations of HAc, con- tions and therefore it disproportionates to afford a sort of dianionic
sists in the fact that the occurrence of peaks IVc–IVa cannot be face-to-face dimeric complex, as proposed in Eq. (7).
explained by any of the species involved in the two-electron and
 
two-proton transfer mechanisms (Eqs. (3)–(6) and/or (3), (4), (5), 2QH− → Q2− , QH2 (7)
(6)). Nevertheless, these results suggest that the species QH− must
be related in some way to that occurring in peak IVc–IVa. There- The proposed complex can be viewed as a dianionic quinhy-
fore, to obtain information about the electrochemical behaviour of drone, which must involve two hydrogen bonding interactions, but
QH− , but in a more simplified experimental system, the product of stronger than in the case of neutral quinhydrone. We have assumed
half-deprotonation of QH2 with tetrabutylammonium hydroxide [Q2− · · ·QH2 ] as notation for this complex; however, the structure
was analysed. should be rather that corresponding to a low barrier proton transfer
200 P.D. Astudillo et al. / Electrochimica Acta 81 (2012) 197–204

_ _ _ _
O H O O H O O H O O H O 2

_O H O _O H O _O H O O H O
_ _ _ _ _ _
[Q 2 ... QH ]
2
[QH ... QH ] [QH ... QH ] [Q ... H,H ... Q] 2

Scheme 1. Resonant structures of the quinhydrone-type dianionic complex.

complex, which can be represented by several resonant structures, pure deuterated acetonitrile, which presents the signals of water
as shown in Scheme 1. and acetonitrile at high field and the signals for hydroquinone at
These strong interactions are commonly found when a proton is 6.63 ppm (C H) and 6.46 ppm (O H). The first signal corresponds to
shared between two equivalent electronegative atoms, where one the aromatic protons and the second one to both hydroxyl protons.
of them is negatively charged [62]. Due to the electrostatic nature In the case of the product of half-neutralization of QH2 with the
of the interaction, it can also be proposed that the occurrence of base (Fig. 3B), the signals for water and deuterated acetonitrile were
the disproportionation process of QH− (Eq. (7)) is possible because also observed, as well as the signals related to the ethyl group of the
of the complex is stabilized through two strong hydrogen bond- tetraethylammonium cation, that is, the triplet and quartet corre-
ing interactions. Although more details about the structure of this sponding respectively to the methyl (CH3 1.19 ppm) and methylene
complex will be discussed later from the point of view of electronic groups (CH2 3.13 ppm). In the same spectra, the signal appearing at
structure calculations, it can be mentioned that an analogue exam- 6.58 ppm is assigned to the aromatic protons in the complex due to
ple of face-to-face dimeric complex, but positively charged, was the fact that it is very close to that found for pure QH2 (6.63 ppm).
recently reported by Smith and coworkers for the case of proto- The presence of this simple signal at 6.58 ppm and the absence of
nated phenylendiamines [63]. any signal corresponding to the O H protons, indicates that the
The proposal of the quinhydrone-type dianionic complex is then product of half deprotonation should be symmetrical. Additionally,
useful to explain the behaviour of the half-neutralization product the fact that no signal was found for these protons even at very low
of QH2 . As shown in the voltammogram of Fig. 2B, the complex field (up to 25 ppm), suggests also that these protons participate
is oxidized at the level of peak IVa, to afford a transient associated in an equilibrium involving a strong hydrogen bonding interaction
radical anion [Q•− · · ·QH2 ], which is dissociated into the free species [62]. In agreement with the electrochemical results, the 1 H NMR
Q•− and QH2 (Eqs. (8) and (9)). In this way, the radical anion Q•− is study suggests also that the product of half-deprotonation of hydro-
oxidized at peak Ia while QH2 is oxidized at peak IIIa. It is remark- quinone is the dianionic quinhydrone linked by strong hydrogen
able that the current intensity of IIIa in Fig. 2B is approximately the bonds in which each hydrogen atom is in fast exchange between
half of that observed in Fig. 2A, which is in agreement with the fact two oxygen atoms, as depicted in Scheme 1.
that the complex was prepared from two equivalents of QH2 while
only one equivalent of this one is released from the oxidation of
complex at peak IVa. 3.3. The dimeric nature of complex by diffusion coefficient
 2−
 −
 •−  measurements
Q , QH2 − e  Q , QH2 (8)
 •−  •− The dimeric nature of the product of half-deprotonation of
Q , QH2  Q + QH2 (9)
hydroquinone can be evidenced considering that the diffusion coef-
Several studies of association of semiquinone and quinone dian- ficient of complex must be smaller than that of the hypothetical
ion with weak proton donors have led to conclude that hydrogen species QH− . Thus, an estimation of the molecular weight of species
bonding interactions with the dianion are stronger than in the case [Q2− · · ·QH2 ] giving rise to peak IVa has been carried out through dif-
of semiquinone [44–46], which supports the proposal that the rad- fusion coefficient measurements. Such estimation in acetonitrile is
ical anion complex [Q•− · · ·QH2 ] is dissociated according to Eq. (9). possible by using a previously reported linear correlation between
As shown in the voltammogram of Fig. 2B, where the potential scan the diffusion coefficient and the molecular weight for a family of
is inverted at 0 V vs SCE, the chemical reversibility of waves Ia–Ic quinones, nitroaromatic compounds, aromatic compounds and fer-
indicates that dissociation of [Q•− · · ·QH2 ] is indeed highly favoured. rocene derivatives (Fig. 4) [64]. The use of such correlation in our
Considering that the position of the peaks of quinones can be shifted case is reliable because the molecular density of the target complex
towards less negative values by effect of hydrogen bonding inter- must be comparable to the average molecular density found for the
actions, the invariance in the position of waves Ia–Ic in Fig. 2B with family of compounds used to construct the mentioned correlation.
respect to those of Fig. 2C, in which the free species Q•− and Q2− Single potential step chronoamperometry has been used to
are involved, suggests that the association equilibrium constant determine the diffusion coefficient of complex [Q2− · · ·QH2 ], tak-
between Q•− and QH2 must be in fact very small. ing as reference the experiment depicted in Fig. 2A. Owing to the
In order to obtain chemical information about the species giv- fact that peak IVa is too close to peak Ia, we decided to apply a
ing rise to the broad peak IVa of Figs. 1 and 2, the product of potential step at −0.265 V vs SCE, where pure diffusion conditions
half-deprotonation of hydroquinone was studied by 1 H NMR spec- are established. According to the voltammogram of Fig. 2B, it has
troscopy in deuterated acetonitrile (Fig. 3). been considered that one electron is consumed for the oxidation
Considering that low contents of water does not modify of [Q2− · · ·QH2 ] while an additional electron is consumed for the
the electrochemical behaviour in a relevant way, as observed oxidation of semiquinone Q•− . Therefore, the calculation of the
in Figs. 1 and 2, the hydroquinone was directly deprotonated diffusion coefficient from the Cottrell behaviour was performed
with tetraethylammonium hydroxide before the 1 H NMR spectra assuming that the electron number is 2. Upon these considera-
recording. Due to the fact that the voltammetric behaviour of the tions, the diffusion coefficient of [Q2− · · ·QH2 ] was determined to
product of half-neutralization of QH2 with base is identical for solu- be (1.97 ± 0.07) × 10−5 cm2 s−1 .
tions 2 mM and 8 mM, the 1 H NMR experiments were performed Fig. 4 shows the linear correlation between the diffu-
at this last concentration. Fig. 3A shows the spectrum of QH2 in sion coefficient and the molecular weight that we have
P.D. Astudillo et al. / Electrochimica Acta 81 (2012) 197–204 201

Fig. 3. 1 H NMR spectrum in deuterated acetonitrile. (A) QH2 8 mM in pure deuterated acetonitrile, (B) Product of half-neutralization of QH2 8 mM with tetraethylammonium
hydroxide.

previously reported [64], whose fitting equation is given by considered to be the result of disproportionation of the anion QH− ,
D[cm2 s−1 ] = 3.52 × 10−5 − 6.59 × 10−8 (Mw). The use of this equa- such as it was depicted by Eq. (7).
tion and the experimental diffusion coefficient have permitted to
extrapolate a value of the molecular weight of complex [Q2− , QH2 ],
which was about MW = 235.5. This value is very close to the the- 3.4. The role of solvent on the complex formation
oretical molecular weight of the proposed complex (MW = 218.2).
This comparison represents additional evidence supporting the The electrochemical oxidation of hydroquinone is sensitive to
proposal of a dianionic-type quinhydrone dimeric complex for the nature of solvents such as acetonitrile and dimethylsulfoxide,
the product of half-deprotonation of hydroquinone, which was whose hydrogen bonding acceptor properties influence the rate of
proton cleavage involved in the reaction mechanism [49]. In this
way, it is interesting to investigate the effect of DMSO on the forma-
tion and stability of the quinone dianion–hydroquinone complex.
Fig. 5 shows the voltammograms obtained in acetonitrile and
DMSO for the proposed complex, which is the half-deprotonation
product of hydroquinone. By comparing Fig. 5A and B, it is observed
that the general features of both voltammogramms are essentially
the same, suggesting that the complex is also formed in DMSO
and it follows an oxidation mechanism analogue to that found in
acetonitrile, giving rise to the formation of semiquinone Q•− and
hydroquinone QH2 (Eqs. (8) and (9)). However, some variations in
the position of peaks are observed. For example, the oxidation peak
IIIa is shifted towards less positive values in DMSO because this sol-
vent increases the deprotonation of the radical cation QH2 •+ formed
during the first electron transfer step of the oxidation mechanism
of QH2 [49]. Concerning peaks Ia–Ic, they are slightly less negative
in DMSO, an effect that has been attributed to stabilization of Q•−
by effect of the larger solvation capacity of DMSO with respect to
acetonitrile [65].
Concerning the broad peaks IVa and IVc, in DMSO, they were
more negative (Epa = −0.746, Epc = −0.994 V vs SCE) than in the
Fig. 4. Plot of the diffusion coefficient of quinones, aromatic hydrocarbons, aro-
matic nitrocompounds and ferrocene derivatives, in acetonitrile +0.1 M n-Bu4 NPF6 case of acetonitrile (Epa = −0.626, Epc = −0.870 V vs SCE). The shift
at 25 ◦ C, against its respective molecular weight [64]. of the peak potential IVa towards more negative values can be
202 P.D. Astudillo et al. / Electrochimica Acta 81 (2012) 197–204

Fig. 5. Cyclic voltammetry of hydroquinone 2 mM in the presence of tetrabuty-


lammonium hydroxide 2 mM, in acetonitrile (A) and dimethylsulfoxide (B), +0.2 M Fig. 6. Cyclic voltammetry of 1,4-benzoquinone 2 mM in the presence of different
n-Bu4 NPF6 , on glassy carbon electrode (3 mm ␾) at 0.1 Vs−1 . concentrations of hydroquinone: (black) 0 mM, (red) 1 mM, (green) 2 mM, (blue)
4 mM; on glassy carbon electrode (3 mm ␾), in DMSO + 0.1 M n-Bu4 NPF6 at 0.1 Vs−1 .
(For interpretation of the references to colour in this figure legend, the reader is
explained by considering that semiquinone Q•− and hydroquinone referred to the web version of the article.)
QH2 are both stabilized, the former by solvation with DMSO,
and QH2 through interaction with this hydrogen bonding accep-
tor solvent. On the contrary, in the case of peak IVc, this signal Q2− and QH2 is qualitatively manifested by the fact that intensity
is the result of the reduction of semiquinone to Q2− coupled of peaks IIc–IIa diminishes at the same time that peaks IVc–IVa
with strong association equilibrium with QH2 . The peak poten- increase their current intensity, such as it was verified in a previous
tial IVc is more negative in DMSO than in the case of acetonitrile study about the chloranil reduction in presence of several amides
because the hydrogen bonding interaction between QH2 and DMSO [66]. As it was previously mentioned, the reason of this strong inter-
is unfavourable for the quinone dianion–hydroquinone complex action must be related to the formation of two hydrogen bonds in
formation. This result suggests the possibility to generate the the ionic-neutral mode. The nature of such interaction is discussed
quinone dianion–hydroquinone complex during the reduction of in the next section from the point of view of electronic structure cal-
1,4-benzoquinone in the presence of hydroquinone. Since the broad culations. In this framework and considering the species depicted in
peaks IVa–IVc are more separated from the quinone–semiquinone Scheme 1 the interconversion between the complexes [Q2− · · ·QH2 ]
peaks in DMSO, this solvent was used for such experiments. and [QH− · · ·QH− ] was considered.

3.5. Complex formation during reduction of benzoquinone in 3.6. Electronic structure calculations of the complexes
presence of hydroquinone [Q2− · · ·QH2 ] and [QH− · · ·QH− ]

It should be noticed that, even when using a strong hydrogen In order to analyse the stability of the quinone
bonding coordinating solvent such as DMSO, the presence of sig- dianion–hydroquinone complex [Q2− · · ·QH2 ] and hydroquinone
nals IVa–IVc indicates that the intermolecular interaction between monoanion dimer [QH− · · ·QH− ], the electronic structure cal-
QH2 and Q2− must be very strong. This effect resulted in the for- culations using Density Functional Theory were performed for
mation of those signals during voltammetric titration experiments the systems studied. For this purpose, four different function-
of 1,4-benzoquinone (Q) with hydroquinone (QH2 ) (Fig. 6). The als were employed, specially to take into account the effect of
voltammetric behaviour of 1,4-benzoquinone (Q) as a function of intermolecular hydrogen-bonding [67,68] and ␲–␲ interactions
several concentrations of hydroquinone (QH2 ) was carried out in within the structures proposed [69]. Full geometry optimization
DMSO. of the structures for both complexes indicates the stability of a
Fig. 6 (black curve) shows the peaks Ic–Ia and IIc–IIa, which face-to-face dimer, as suggested above (Fig. 7).
represent the classical behaviour of benzoquinone 2 mM in pure A collection of selected results from the geometric and energetic
DMSO. In the presence of 1 mM QH2 (red curve), signals IIc–IIa analysis of the complexes is presented in Table 1. From the results
decrease in intensity while the broad signal IVc–IVa appears at sim- obtained, it is apparent that the use of the MPW1PW1 functional
ilar potential values as the broad signal IVc–IVa observed in Fig. 5B. leads to the lowest values of total energies for the series of opti-
This effect is magnified by increasing the concentration of QH2 up mized structures. It should be noticed that this functional has been
to 2 mM (green curve). Even at this concentration conditions (1:1 used to analyse stability effects in dimers where ␲–␲ interactions
relationship), a small fraction of the signals corresponding to the play a dominant role such as in benzene dimers [69]. However,
free dianion Q2− (peaks IIc–IIa) are still observed, implying that the there is a slight shift between the planes of both aromatic rings
formation of the complex between Q2− and QH2 is not 100% quan- (Fig. 7, upper structures), and this shift diminishes the interaction
titative. Upon increasing even more the concentration of QH2 until between the ␲ electrons of the rings. Therefore, most of the stabil-
4 mM (blue curve), only the signals Ic–Ia and IVc–IVa are mainly ity effects should be addressed to the hydrogen bonding effects of
observed (Fig. 6B). This result confirms that the dianionic type quin- the OH terminal groups. In particular, the use of BHandHLYP func-
hydrone complex can also be formed by direct interaction between tional led to a significant low value of the HOMO level, compared
Q2− and QH2 . The high value for the association constant between with the MPW1PW1 and PBE functional, the latter being known as
P.D. Astudillo et al. / Electrochimica Acta 81 (2012) 197–204 203

Table 1
Selected geometric parameters and energies of the dimeric structures considered in
this work.

Dimer Etotal /kJ mol−1 EHOMO /eV r O H/nm r O· · ·H/nm


− −
[QH · · ·QH ] −2,007,358a 2.4847 0.09884 0.18808
−2,006,647b 2.5045 0.09894 0.18717
−2,006,676c 1.6904 0.09722 0.19895
−2,006,471d 0.2727 0.09886 0.18793
[Q2− · · ·QH2 ] −2,007,325a 3.0882 0.10556 0.15045
−2,006,617b 3.0831 0.10598 0.14910
−2,006,638c 2.5587 0.10156 0.16205
−2,006,437d 0.9099 0.10545 0.15106

Each line shows results from the different functionals employed.


a
MPW1PW1.
b
MPW1PBE.
c
BHandHLYP.
d
Long range corrected PBE.

that the PBE electron correlation parameters are required for a


correct estimation of the electronic effects of this type of bond.
Considering the possibility of the fast interchange between both
Fig. 7. Optimized structures of (A) [QH− · · ·QH− ] and (B) [Q2− · · ·QH2 ] at the LC- [QH− · · ·QH− ] and [Q2− · · ·QH2 ] structures suggested through the
PBE/6-311 + +G(d,p) level. Upper figures represent views from H O· · ·H sides; lower experimental study, an analysis of a potential energy surface (PES)
figures show lateral views of both dimers.
was performed by changing the distances between the OH bond
and the hydrogen bonding between the O− residue and the neigh-
bouring H atoms. Fig. 8 shows the contour lines of the constructed
useful for accounting of hydrogen bonding effects. This effect is not PES with the LC-PBE functional along with the plot of the mini-
associated to significant changes in geometric parameters for the mum energies of the PES, from which activation parameters were
optimized structures of both dimers, but suggests that increasing calculated.
the amount of Hartree–Fock exchange is relevant to indicate a more With this analysis, a transition state can be identified, from
stable structure, as expected for this type of polar interactions. This which activation parameters for the interconversion between
correction effect has been addressed before in the analysis of reac- structures [QH− · · ·QH− ] and [Q2− · · ·QH2 ] were estimated. For
tive trends in substituted quinone systems [28], though this was instance, an activation energy of 108.4 kJ mol−1 was determined for
relevant in single molecules. As this type of effects diminishes sig- converting structure [QH− · · ·QH− ] into [Q2− · · ·QH2 ], meanwhile,
nificantly with the distance between the considered species, long for the opposite process, an activation energy of 28.4 kJ mol−1 was
range corrections are necessary to account for this effect in the found. Even though these theoretical results suggest that struc-
total energy. For this purpose, the long range corrected functional ture [QH− · · ·QH− ] is more stable than [Q2− · · ·QH2 ], the calculated
PBE (LC-PBE) was also employed, showing a significant decrease activation energies are lower than the thermal activation factor at
in the HOMO value for both [QH− · · ·QH− ] and [Q2− · · ·QH2 ] com- 298.15 K (NA kB T ∼ 2480.2 kJ mol−1 ); thus, it is possible than with
plexes (Table 1). It should also be noticed that hydrogen bonding the temperature at which the experimental data were obtained,
distances calculated with this functional are comparable with those rapid thermal interconversion between both structures is occur-
obtained with the hybrid MPW1PBE functional, thus suggesting ring.

Fig. 8. (A) Potential energy surface of the transition between structures [QH− · · ·QH− ] and [Q2− · · ·QH2 ] at the LC-PBE/6-311 + +G(d,p) level. (B) Variation of the reaction energy
along the minima of PES. In both figures, the position of conformers [QH− · · ·QH− ] and [Q2− · · ·QH2 ] are indicated.
204 P.D. Astudillo et al. / Electrochimica Acta 81 (2012) 197–204

4. Conclusions [26] B.R. Eggins, J.Q. Chambers, Journal of the Electrochemical Society 11 (1970)
186.
[27] P. Zuman, Substituent Effects in Organic Polarography, Plenum Press, NY, 1967.
The electrochemistry of the 1,4-benzoquinone and hydro- [28] C. Frontana, A. Vazquez-Mayagoitia, J. Garza, R. Vargas, I. González, Journal of
quinone system in organic media was analysed. Using cyclic Physical Chemistry A 110 (2006) 9411.
voltammetry, it was found that the electrochemical behaviour of [29] M. Aguilar-Martinez, G. Cuevas, M. Jimenez-Estrada, I. González, B. Lotina-
Hennsen, N. Macias-Ruvalcaba, Journal of Organic Chemistry 64 (1999) 3684.
quinone in the presence of small amounts of proton donors can- [30] M. Gómez, F.J. González, I. González, Journal of Electroanalytical Chemistry 578
not be explained in terms of the classical mechanisms of quinone (2005) 193.
reduction, due to the fact that the anion QH− disproportionates into [31] C. Frontana, I. Gonzalez, Journal of the Brazilian Chemical Society 16 (2005)
299.
a face-to-face dianionic quinhydrone associated complex stabilized
[32] N.A. Macias, D.H. Evans, Journal of Physical Chemistry C 114 (2010) 1285.
by strong intermolecular hydrogen bonds. It was demonstrated [33] L.S. Hernández-Muñoz, M. Gómez, F.J. González, I. González, C. Frontana,
that this complex can be formed by the reduction of benzoquinone Organic & Biomolecular Chemistry 7 (2009) 1896.
[34] M. Shamsipur, A. Siroueinejad, B. Hemmateenejad, A. Abbaspour, H. Sharghi, K.
in either the presence of stoichiometric amounts of weak proton
Alizadeh, S. Arshadi, Journal of Electroanalytical Chemistry 600 (2007) 345.
donors or by half-deprotonation of hydroquinone, as also evaluated [35] C. Ruessel, W. Jaenicke, Journal of Electroanalytical Chemistry 199 (1986) 139.
by 1 H NMR experiments. Theoretical analysis of energetic barriers [36] B. Uno, N. Okumura, M. Goto, K. Kano, Journal of Organic Chemistry 65 (2000)
for hydrogen transfers involved in the complexes formation indi- 1448.
[37] N. Okumura, B. Uno, Bulletin of the Chemical Society of Japan 72 (1999) 1213.
cated that interconversion between [QH− · · ·QH− ] and [Q2− · · ·QH2 ] [38] B.R. Clark, D.H. Evans, Journal of Electroanalytical Chemistry 69 (1976) 181.
species is possible, and that it could be driven by thermal effects [39] C. Frontana, B. Frontana-Uribe, I. González, Journal of Electroanalytical Chem-
rather than thermodynamic stability. istry 573 (2004) 307.
[40] S. Klod, L. Dunsch, Magnetic Resonance in Chemistry 49 (2011) 725.
[41] T. Hoshino, M. Oyama, S. Okazaki, Journal of Electroanalytical Chemistry 472
Acknowledgements (1999) 91.
[42] H. Park, M. Oyama, Electroanalysis 14 (2002) 1269.
[43] D.H. Evans, D.A. Griffith, Journal of Electroanalytical Chemistry 136 (1982) 149.
P.D. Astudillo and D. Valencia acknowledge CONACyT for [44] N. Gupta, H. Linschitz, Journal of the American Chemical Society 119 (1997)
their Ph.D. scholarships. M.A. González-Fuentes also acknowl- 6384.
edges CONACyT for a postdoctoral fellowship. C. Frontana thanks [45] B. Uno, A. Kawabata, K. Kano, Chemistry Letters 6 (1992) 1017.
[46] M. Gómez, F.J. González, I. Gonzalez, Journal of the Electrochemical Society 150
CONACyT-México for support through project number 107037 (2003) E527.
(Fondo de Investigación Científica Básica SEP-CONACyT 2008). [47] S. Ahmed, A.Y. Khan, R. Qureshi, M.S. Subhani, Russian Journal of Electrochem-
istry 43 (2007) 811.
[48] B. Uno, N. Okumura, K. Seto, Journal of Physical Chemistry A 104 (2000) 3064.
References [49] P.D. Astudillo, J. Tiburcio, F.J. González, Journal of Electroanalytical Chemistry
604 (2007) 57.
[1] J.L. Cape, M.K. Bowman, D.M. Kramer, Phytochemistry 67 (2006) 1781. [50] C. Costentin, Chemical Reviews 108 (2008) 2145.
[2] J.E.Q. Walker, Quarterly Reviews of Biophysics 25 (1992) 253. [51] M. Salas, M. Gómez, F.J. González, B. Gordillo, Journal of Electroanalytical Chem-
[3] E.A. Hillard, F.C. de Abreu, F. Caxico, D.C. Ferreira, G. Melo, M.O.F. Jaouen, C. istry 543 (2003) 73.
Goulart, Amatore, Chemical Communications 23 (2008) 2612. [52] F. D’Souza, G.R. Deviprasad, Journal of Organic Chemistry 66 (2001) 4601.
[4] J.Q. Chambers, in: S. Patai (Ed.), Chemistry of Quinonoid Compounds, vol. 2, Part [53] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,
1, 1988, pp. 719–757. G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato,
[5] M. Kurosi, E. Begari, Molecules 15 (2010) 1531. X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M.
[6] J. Oldenburg, M. Marinova, C. Muller-Reible, M. Watzka, Vitamins & Hormones Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y.
78 (2008) 35. Honda, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery Jr., J.E. Peralta, F. Ogliaro,
[7] J.E. Sadler, Nature 427 (2004) 493. M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, T. Keith, R.
[8] J. Hirst, Biochemical Journal 425 (2010) 327. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J.
[9] A.D. Vinogradov, Biochimica et Biophysica Acta 1777 (2008) 729. Tomasi, M. Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken,
[10] M. Sarasate, Science 283 (1999) 1488. C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R.
[11] M.R. Gunner, J. Madeo, Z. Zhu, Journal of Bioenergetics and Biomembranes 40 Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski,
(2008) 509. G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, O. Farkas, J.B.
[12] P. Joliot, A. Joliot, G. Johnson, Advances in Photosynthesis and Respiration 24 Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian 09 Revision B.01, Gaussian,
(2006) 639. Inc, Wallingford, CT, 2010.
[13] U. Siggel, Bioelectrochemistry and Bioenergetics 3 (1976) 302. [54] Used as defined in Gaussian 09 Manual: 0.5EXHF + 0.5EXLSDA + 0.5EXBecke88
[14] M. Hervás, J.A. Navarro, M.A. de la Rosa, Accounts of Chemical Research 36 + ECLYP.
(2003) 798. [55] C. Adamo, V. Barone, Journal of Chemical Physics 108 (1998) 664.
[15] V.K. Tandon, R.V. Singh, D.B. Yadav, Bioorganic & Medicinal Chemistry Letters [56] A.D. Becke, Journal of Chemical Physics 104 (1996) 1040.
14 (2004) 2901. [57] Y. Tawada, T. Tsuneda, S. Yanagisawa, T. Yanai, K. Hirao, Journal of Chemical
[16] S. Hayashi, H. Ueki, Y. Ueki, H. Aoki, K. Tanaka, J. Fujimoto, K. Katsukawa, M. Physics 120 (2004) 8425.
Mori, Chemical & Pharmaceutical Bulletin 11 (1963) 948. [58] O.A. Vydrov, G.E. Scuseria, Journal of Chemical Physics 125 (2006) 234109.
[17] D.N. Singh, N. Verma, S. Raghuwanshi, P.K. Shukla, D.K. Kulshreshtha, Bioor- [59] O.A. Vydrov, J. Heyd, A. Krukau, G.E. Scuseria, Journal of Chemical Physics 125
ganic & Medicinal Chemistry Letters 16 (2006) 4512. (2006) 074106.
[18] G. Meazza, F.E. Dayan, D.E. Wedge, Journal of Agricultural and Food Chemistry [60] O.A. Vydrov, G.E. Scuseria, J.P. Perdew, Journal of Chemical Physics 126 (2007)
51 (2003) 3824. 154109.
[19] E.N. da Silva, T.T. Guimaraes, R.F.S. Menna-Barreto, M.C.F.R. Pinto, C.A. de [61] C. Amatore, J.-M. Savéant, Journal of Electroanalytical Chemistry 85 (1977) 27.
Simone, C. Pessoa, B.C. Cavalcanti, J.R. Sabino, C.K.Z. Andrade, M.O.F. Goulart, [62] C.L. Perrin, J.B. Nielson, Annual Review of Physical Chemistry 48 (1997)
S.L. de Castro, A.V. Pinto, Bioorganic & Medicinal Chemistry 18 (2010) 3224. 511.
[20] K.A.L. Ribeiro, C. Monteiro de Carvalho, M.T. Molina, E.P. Lima, E. Lopez- [63] L.A. Clare, L.E. Rojas-Sligh, S.M. Maciejewski, K. Kangas, J.E. Woods, L.J. Deiner,
Montero, J.R.M. Reys, M.B. Farias de Oliveira, A.V. Pinto, A.E.G. Santana, M.O.F. A. Cooksy, D.K. Smith, Journal of Physical Chemistry C 114 (2010) 8938.
Goulart, Acta Tropica 111 (2009) 44. [64] D.P. Valencia, F.J. González, Electrochemistry Communications 13 (2011) 129.
[21] E.N. da Silva, C.F. de Deus, B.C. Cavalcanti, C. Pessoa, L.V. Costa-Lotufo, R.C. Mon- [65] J.H. Wilford, M.D. Archer, Journal of Electroanalytical Chemistry 190 (1985)
tenegro, M.O. de Moraes, M.C.F.R. Pinto, C.A. de Simone, V.F. Ferreira, M.O.F. 271.
Goulart, C.K.Z. Andrade, A.V. Pinto, Journal of Medicinal Chemistry 53 (2010) [66] M. Gómez, C.Z. Gómez-Castro, I.I. Padilla-Martínez, F.J. Martínez-Martínez, F.J.
504. González, Journal of Electroanalytical Chemistry 567 (2004) 269.
[22] Y. Chen, L. Hu, Medicinal Research Reviews 29 (2009) 29. [67] R. Li, K. Hongwei, F. Gang, X. Xin, Y. Yijing, Journal of Chemical Theory and
[23] S.I. Bailey, I.M. Ritchie, Electrochimica Acta 30 (1985) 3. Computation 5 (2009) 86.
[24] M. Quan, D. Sanchez, M.F. Wasylkiw, D.K. Smith, Journal of the American Chem- [68] N.A. Benedek, K. Latham, I.K. Snook, I. Yarovsky, Journal of Physical Chemistry
ical Society 129 (2007) 12847. B 110 (2006) 19605.
[25] M. Aguilar-Martínez, N.A. Macías-Ruvalcaba, J.A. Bautista-Martínez, M. Gómez, [69] M.J. González-Moa, M. Mandado, R.A. Mosquera, Journal of Physical Chemistry
F.J. González, I. González, Current Organic Chemistry 8 (2004) 1721. A 111 (2007) 1998.

You might also like