You are on page 1of 9

Electrochimica Acta 81 (2012) 292–300

Contents lists available at SciVerse ScienceDirect

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

Understanding the electro-catalytic oxidation mechanism


of urea on nickel electrodes in alkaline medium
Vedasri Vedharathinam, Gerardine G. Botte ∗,1
Center for Electrochemical Engineering Research, Chemical and Biomolecular Engineering Department, 165 Stocker Center, Ohio University, Athens, OH 45701, United States

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

Article history: In this paper, the electrocatalytic behavior of urea oxidation on a nickel electrode in alkaline medium was
Received 5 May 2012 investigated using various electrochemical techniques: cyclic voltammetry (CV), linear sweep voltam-
Received in revised form 2 July 2012 metry (LSV), and rotating disc electrode voltammetry (RDE). Possible mechanisms have been proposed
Accepted 3 July 2012
on the basis of the cyclic voltammetry data. The urea oxidation current shows a strong dependence on the
Available online 14 July 2012
concentration of urea and potassium hydroxide that enables the experimental determination, by Tafel
studies, of the kinetic reaction orders for the oxidation of urea. The Tafel studies show reaction orders
Keywords:
of 0.3 and 2, with respect to urea and OH− concentrations, respectively. The diffusion coefficient of urea
Urea electro-oxidation
Nickel electrocatalyst
was found to be 0.85 × 10−5 cm2 s−1 . The presence of a mixed-controlled process (diffusion and kinetics)
Hydrogen production was confirmed for the oxidation of urea.
Nickel oxyhydroxide © 2012 Elsevier Ltd. All rights reserved.
Fuel cells

1. Introduction at the cathode that can be collected as a valuable fuel and clean
water is obtained as a by-product [1,4]. This technology can be
To meet the increasing global demands for energy, alternative a low-cost substitute when coupled with wastewater treatment
clean energy sources independent of fossil fuels, must be devel- plants due to the capital expensive current de-nitrification pro-
oped. The use of abundantly available wastewater for hydrogen cesses.
production has been attaining recent attention as an alternative Nobel metal catalysts like Pt, Ir, Ru, etc., are expensive and hence
source of energy [1–11]. Urea rich wastewater has been identi- inexpensive non-noble metal catalysts are being investigated for
fied as a good source for hydrogen production in alkaline medium the oxidation of organic molecules in alkaline medium [1,4,16,17].
[1–4,6,12]. The major constituent of human or animal waste on Nickel and nickel-based modified electrodes are extensively used
earth is urine, containing about 2–2.5 wt.% of urea [13] suggest- as electrocatalysts in the oxidation and determination of organ-
ing the availability of a considerable amount of urea in municipal ics such as methanol [18], ethanol [19], glucose [20], cyclohexanol
wastewater [14]. Also, a large amount of wastewater with varying [21], aspirin [22], etc. Botte [2] and Boggs et al. [1] had successfully
concentrations of urea is produced during the industrial synthesis demonstrated the electrolysis of urea to produce hydrogen using Ni
of urea. The amount of urea and its derivatives in our environment, electrodes in alkaline medium. Some studies have been done in the
especially in groundwater is increasing and its natural degrada- past years to identify the mechanism of urea hydrolysis in the pres-
tion to nitrate and nitrite by microorganisms leads to high health ence of urease (with Ni active sites), both experimentally [23,24]
risks in humans [15]. The available de-nitrification technologies are and theoretically [25,26]. But the research investigating the electro-
very expensive and inefficient and hence a proficient methodology chemical oxidation mechanism of urea on nickel-based electrodes
should be devised. in alkaline medium is limited. Daramola et al. [25] studied the disso-
Urea electrolysis would be an efficient way for hydrogen pro- ciation rates of urea in the presence of nickel oxyhydroxide (NiOOH)
duction from urea rich wastewater thus potentially using urine using density functional theory (DFT) methods and reported the
[1,2,4], the product of human/animal excretion as an energy source, desorption of CO2 from urea as the rate-limiting step. They also
as shown in Fig. 1. Urea is electrochemically oxidized at the reported that the NiOOH catalyst surface could be deactivated by
anode producing N2 and CO2 , whereas pure hydrogen is evolved the surface blockage due to CO groups.
In previous studies, the authors have demonstrated that human
urine with an average concentration of 0.33 M urea can be elec-
∗ Corresponding author. Tel.: +1 740 593 9670; fax: +1 740 593 0873. trochemically oxidized to a non-toxic N2 product with the aid of
E-mail address: botte@ohio.edu (G.G. Botte). a low-cost transition metal catalyst (nickel) in alkaline medium
1
ISE member. according to the following reactions [1]:

0013-4686/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.electacta.2012.07.007
V. Vedharathinam, G.G. Botte / Electrochimica Acta 81 (2012) 292–300 293

250

200

150

i/mA cm-2
5 M KOH + 0.33 M urea
100

Fig. 1. Schematic representation of urea electrolysis unit and its applications. The 50
electrochemical conversion of urea rich wastewater can produce N2 and CO2 at the
anode and H2 at the cathode, thus producing clean water as a byproduct. The pure
H2 thus obtained can be used as fuel for fuel cells to produce electricity.
0 5 M KOH
Anode:
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Ni(OH)2(s) + OH−  NiOOH(s) + H2 O(l) + e− (1)
E/V (Hg/HgO)
CO(NH2 )2(aq) + 6OH− → N2(g) + 5H2 O(l) + CO2(g) + 6e− (2)
Fig. 2. Cyclic voltammogram of nickel electrode obtained in 5 M KOH in the absence,
Cathode: and presence of 0.33 M urea at a scan rate of 10 mV s−1 . A pair of redox peak cor-
responding to Ni(OH)2 and NiOOH is noticed during the forward and reverse scan,
6H2 O(l) + 6e− → 3H2(g) + 6OH− (3) respectively in the absence of urea. In the presence of urea, the anodic current den-
sity increases drastically at potentials higher than 0.35 V vs. Hg/HgO representing a
Overall: strong electrocatalytic activity of the nickel electrode toward the oxidation of urea.

CO(NH2 )2(aq) + H2 O(l) → N2(g) + 3H2(g) + CO2(g) (4)

It has been suggested that Ni2+ undergoes oxidation to its active electrodeposition of nickel. The substrate was mounted on a Teflon
Ni3+ form, as per Eq. (1) that catalyzes the electrochemical oxida- shaft with an exposed surface area of 0.126 cm2 and immersed in
tion reaction of urea; however, the mechanism of the urea oxidation a Watt’s bath solution consisting of 280 g L−1 NiSO4 ·6H2 O, 40 g L−1
reaction is still unclear. The purpose of this paper is to gain a NiCl2 ·6H2 O, and 30 g L−1 H3 BO4 solvated with ultrapure water (Alfa
better understanding of the electrochemical oxidation of urea on Aesar, HPLC grade). The plating solution was stirred at 60 rpm using
nickel in alkaline medium. Different electrochemical techniques a 25.4 mm × 9.5 mm magnetic stir bar at 45 ◦ C in a 100 mL beaker
such as cyclic voltammetry (CV), linear sweep voltammetry (LSV), during electrodeposition. A Solartron 1281 multiplexer potentio-
and rotating disc electrode (RDE) voltammetry were used to inves- stat was used to deposit Ni with a loading of 1.50 ± 0.20 mg cm−2
tigate the mechanism and identify the kinetic parameters of the potentiostatically at −0.80 V vs. Ag/AgCl (saturated KCl) refer-
urea oxidation reaction. ence electrode equipped with a Luggin capillary and a platinum
foil counter electrode (Sigma–Aldrich, 0.05 mm thick 99.99%). The
2. Experimental methods nickel loading was chosen to have a sufficient amount of nickel
on the substrate to catalyze the urea oxidation reaction. The elec-
2.1. Reagents and experimental setup trodeposited nickel electrode was then disconnected, rinsed with
ultrapure water, and used in different experiments.
All chemicals and supplies used were of high purity (>99.9%)
and analytical grade supplied from Fisher Scientific. Ultra-pure 3. Results and discussion
water (Alfa Aesar, HPLC grade) was used throughout, unless oth-
erwise specified. The electrochemical tests were carried out using 3.1. Electrocatalytic oxidation of urea at the nickel electrode
a Solartron 1281 multiplexer potentiostat in a conventional three-
electrode cell setup with a nickel anode (working electrode), a Pt 3.1.1. Cyclic voltammetry
cathode (counter electrode), and an Hg/HgO reference electrode. Fig. 2 shows a typical cyclic voltammogram (CV) of the nickel
The reference electrode was supported using a Luggin capillary electrode in the absence and presence of 0.33 M urea in 5 M KOH
filled with the respective supporting electrolyte (KOH) used in solution at a scan rate of 10 mV s−1 . A pair of redox peak appears
the experiment. The fifth pseudo-steady state voltammograms at 387 mV and 260 mV in the anodic and cathodic regions, respec-
(sustained periodic state) are reported in the paper. All the voltam- tively in the absence of urea that corresponds to the maximum
mograms reported in this study were performed between 0 and currents for the Ni2+ /Ni3+ redox couple according to reaction (1)
0.8 V vs. Hg/HgO as per the requirements of the respective experi- [1]. The anodic peak corresponds to the oxidation of Ni(OH)2 (Ni2+ )
ment. to NiOOH (Ni3+ ) whereas the cathodic peak is assigned to the
reduction of NiOOH (Ni3+ ) to Ni(OH)2 (Ni2+ ). During the forward
2.2. Anode preparation scan, the oxidation of Ni to Ni(OH)2 in alkaline media may usually
form two different crystallographic phases [27–30] i.e. ␣-Ni(OH)2
The anode was prepared by machining a titanium (inert) rod and ␤-Ni(OH)2 which are hydrous and anhydrous forms, respec-
(Sigma–Aldrich, 0.7 cm dia.) to 0.5 cm dia. and 1 cm long cylinders. tively. With time and increase in potentials, the unstable ␣-Ni(OH)2
The surface was roughened using a sandblaster (Crystal Mark sand- transforms to a less hydrated and more stable ␤-Ni(OH)2 phase
blaster, 27.5 ␮m aluminum oxide powder) under dry conditions at as reported by Desilvestro et al. [27]. At more positive potentials,
60 psi, followed by sonication in a solution containing 1:1 ratio of ␤-Ni(OH)2 is oxidized to ␤-NiOOH which is partially converted to ␥-
water and acetone, respectively for 10 min. The substrate was fur- NiOOH and accumulates on the surface [31,32]. The cathodic peak
ther rinsed thoroughly with deionized water and dried prior to the that appeared during the reverse scan indicates the presence of a
294 V. Vedharathinam, G.G. Botte / Electrochimica Acta 81 (2012) 292–300

different phase of nickel oxide probably due to the reduction of 3.1.2. Effect of scan rate on the catalytic currents
␥-NiOOH to ␣ or ␤-Ni(OH)2 [27,31]. The effect of scan rate on the electrocatalytic oxidation of urea on
The presence of urea in KOH solution leads to an increase in the a nickel electrode was investigated further and reported in Fig. 3a.
anodic current density at potentials higher than 0.35 V vs. Hg/HgO The anodic current density increases rapidly with increasing the
representing a strong electrocatalytic activity of the nickel elec- potential scan rate. Also, a shift in the anodic peak potential to pos-
trode toward the oxidation of urea. An anodic peak appears at 0.46 V itive values indicates a change in the reaction kinetics between urea
vs. Hg/HgO followed by a decrease in the current density and the and Ni3+ at higher scan rates. The specific adsorption of urea on
oxygen evolution reaction occurs at higher potentials. It is visible Ni3+ sites can be considered as a reason for the shift of its oxidation
that the oxygen evolution reaction (OER) happens at more positive potential toward more positive values [41]. The linear dependency
potentials when urea is present in the solution (see Fig. 2) which of the anodic peak current (Ipa ) with the square root of the scan rates
may be due to the higher affinity of urea adsorption on Ni3+ species (1/2 ) (Fig. 3b) indicates the probability of a diffusion-controlled
than OH− ions that delays the OER [33,34]. Interestingly, another process for the urea oxidation [33]. It is also interesting to note from
anodic peak is noticed at 0.45 V vs. Hg/HgO during the reverse scan Fig. 3c that the anodic peak potential (Ep ) increases linearly with the
which lies in the urea oxidation region (above 0.35 V and below logarithm of the potential scan rate suggesting kinetic limitations
0.52 V vs. Hg/HgO) observed in the forward scan. This behavior can on the urea oxidation [33].
be attributed to the further oxidation of urea molecules due the The voltammetric profile for urea oxidation on the nickel elec-
regeneration of the active sites (Ni2+ /Ni3+ ) on the electrode sur- trode in alkaline medium shows the presence of an irreversible
face which were previously blocked by urea molecules, reaction process. Hence, the diffusion coefficient (D) of urea in KOH solution
products, or intermediates during the forward scan. The adsorbed can be determined using Eq. (6) [41,42]:
molecules are later oxidized to its products due to the still favorable
applied potential [16]. The low anodic current density observed Ipa = 2.99 × 105 n[(1 − ˛)no ]1/2 ACD1/2 1/2 (6)
during the reverse scan compared with the forward scan may indi- where ‘C’ is the concentration of reactant in the bulk solution
cate that the previously blocked active sites for urea oxidation were (C(urea) = 0.33 M), ‘A’ is the geometrical surface area of the nickel
not completely regenerated. Surely, the anodic current density in electrode (A = 0.126 cm2 ), ‘n’ is the total number of electrons
the reverse scan drops as the unfavorable cathodic potentials are involved in the oxidation of urea to its products (n = 6), ‘no ’ is
reached. Also, a considerable decrease in the cathodic charge is the number of electrons involved in the rate-determining step
noticed during the reverse scan. The dramatic increase in the anodic (no = 1), ‘˛’ is the electron transfer coefficient and ‘D’ is the diffusion
peak during the forward scan and the concurrent decrease in the coefficient of reactant in solution. The constant used in the equa-
cathodic peak in the reverse scan are consistent with a behavior tion corresponds with C mol−1 V−1/2 [43]. The value of ‘˛’ can be
typical to the catalytic regeneration effect (EC mechanism). The determined from Fig. 3c using the linear dependency of the peak
principle of the EC mechanism involves the chemical oxidation of potential Epa with the logarithm of the scan rate , according to:
urea to its products along with the reduction of Ni3+ to Ni2+ species  0.03 
which is coupled with the electrochemical regeneration of Ni2+ to Epa = k + log v (7)
Ni3+ , as per the reactions (1) and (5) [35–39]. The regeneration of ˛no
Ni3+ catalyst will result in the loss of reversibility and increase in where ‘k’ is a constant. The value of ˛ was calculated as 0.78 from
height of the Ni3+ /Ni2+ wave upon addition of increasing amounts the slope of Fig. 3c and substituted into Eq. (6) to calculate the
of urea in the KOH solution [37]. The effect of urea concentration on diffusion coefficient of urea. Thus a diffusion coefficient (Durea )
the oxidation reaction will be addressed in Section 3.1.4. As per the value of 0.85 × 10−5 cm2 s−1 is obtained from the voltammograms
EC mechanism, the catalytic enhancement of the Ni2+ /Ni3+ wave (Fig. 3a–c) for the electrochemical oxidation of urea on a nickel elec-
will contain information on the kinetics of the chemical oxidation trode. A plot of the scan rate-normalized current density vs. the scan
of urea (reaction (5)) [37]. rate is shown in Fig. 3d that exhibits the characteristics shape of a
typical catalyst regeneration mechanism (EC ), as in reactions (1)
Catalyst regeneration:
and (5) [36,44]. This result suggests a possible catalyst regenera-
tion mechanism occurring during the oxidation of urea on Ni3+ in
alkaline medium.

3.1.3. Effect of different concentrations of KOH


The effect of KOH concentration on the electro-oxidation of
urea on a nickel electrode at a scan rate of 10 mV s−1 is shown in
Fig. 4. The analysis implies an improvement in the oxidation cur-
rent density with increasing pH. Since OH− activity has a strong
dependence on the formation of NiOOH species that catalyzes the
The products of the electrocatalytic oxidation of urea have been urea oxidation reaction, it is expected to see an increase in the
identified as N2 and CO2 in our previous study [1]. Similar mecha- anodic current density with increasing OH− concentration. Also,
nisms have been reported in the electrochemical oxidation of other the onset potential for the urea oxidation shifts to more negative
organic compounds like methanol [40], ethanol [33], aspirin [22], values with increasing OH− concentration. This explains that the
glucose [20] and cyclohexanol [21] using nickel based catalyst in urea oxidation reaction becomes thermodynamically more favor-
alkaline medium. able at higher KOH concentrations, as expected. The cathodic peak
Also, the decrease in cathodic charge in the presence of urea can that is due to the reduction of NiOOH species also has a strong
be explained by the fact that the applied potential is now being influence on the OH− activity. The inset in Fig. 4 shows a decrease
split between two oxidation reactions i.e. Ni(OH)2 oxidation and in the cathodic current density with increasing OH− concentration.
urea oxidation (according to reactions (1) and (2)). Whereas, in the The result is consistent with the cyclic voltammogram obtained in
absence of urea, the applied potential is used only for the oxida- 5 M KOH and 0.33 M urea (Fig. 2). Also, the charge associated with
tion of Ni(OH)2 to NiOOH. Hence, the split in the applied potential the cathodic peak in the absence and presence of urea in various
between two oxidation reactions can also be a possible explanation KOH concentrations is calculated by integrating the area under the
for the decreased cathodic charge in the presence of urea. respective cathodic peaks and reported in Fig. 5. The plot clearly
V. Vedharathinam, G.G. Botte / Electrochimica Acta 81 (2012) 292–300 295

300 60
(a) (b)

200 45

i/mA cm-2
increasing

Ipa/mA
100 scan rate 30

0 15

-100 0
0.0 0.2 0.4 0.6 0.8 0.1 0.2 0.3 0.4 0.5
E/V (Hg/HgO) ν 1/2/mV sec-1/2

0.54 22
(c) (d)

0.53

ipa ν-1/2/mA cm-2


20
Epa/ V (Hg/HgO)

0.52 18

0.51 16

0.50 14
-1.75 -1.50 -1.25 -1.00 -0.75 4 6 8 10 12 14
log/(ν, V s-1) ν/mV s-1

Fig. 3. Cyclic voltammogram of nickel electrode in 5 M KOH solution containing 0.33 M urea: (a) at scan rates such as 20, 40, 60, 80, 100, 140, and 180 mV s−1 , (b) plot of Ipa
vs. square root of scan rate, (c) plot of Epa vs. logarithm of scan rate, and (d) variation of the scan rate normalized current (Ipa −1/2 ) with scan rate. The anodic current density
increases rapidly with increasing the potential scan rate. Ipa vs. 1/2 indicates a diffusion-controlled process for urea oxidation at the nickel electrode. But, Epa increases
linearly with the potential scan rate suggesting kinetic limitations.

80 60
0
0 1M KOH
Qcathode /mC cm-2

1M KOH + 0.33 M Urea


-1
i/mA cm-2

60 1 55 -10
2
Decrease in Qcathode / %

-2 3
4
-20
-2

7 5 50
i/mA cm

40 -3 6
-30
0.3 0.4 0.5 0.6
E/V (Hg/HgO) 45 0 1 2 3 4 5
20 6 5 CKOH /M
4 3
7 2

1 40
0

0.2 0.3 0.4 0.5 0.6 0.7 0.8 35


0 1 2 3 4 5 6
E/V (Hg/HgO)
CKOH /M
Fig. 4. Effect of KOH concentrations on the electrocatalytic oxidation of urea at
nickel electrode in 0.33 M urea at a scan rate of 10 mV s−1 . The concentration of Fig. 5. Percentage decrease in cathodic charge from CVs of nickel electrode (Fig. 4)
KOH used was (1) 0.25 M, (2) 0.5 M, (3) 1 M, (4) 2 M, (5) 3 M, (6) 4 M, and (7) 5 M. in 0.33 M urea solution at different KOH concentrations at 10 mV s−1 . Inset shows
Inset shows the effect of KOH concentration on the cathodic peak. The CVs imply an the variation in cathodic charge with different concentrations of KOH (data taken
increase in the anodic current density with increasing the pH due to the increased from Fig. 4). It is shown that the decrease in the cathodic charge is linear up to 1 M
amount of NiOOH produced which catalyses the oxidation of urea. The inset shows a KOH exhibiting a diffusion limitation of the OH− species to the catalyst surface, after
decrease in the cathodic current density in the presence of urea with increasing OH− which it deviates from linearity. This deviation may be due to the complete coverage
concentration which can be either due to the indirect electron transfer mechanism of the catalyst surface with OH− ions thus blocking the access of urea molecules for
or the partial distribution of potential for the oxidation of Ni(OH)2 along with the further oxidation. The inset shows the decrease in the cathodic charge is high (50%)
oxidation of urea. at low concentrations of KOH and low (39%) at high concentrations of KOH.
296 V. Vedharathinam, G.G. Botte / Electrochimica Acta 81 (2012) 292–300

(a) 40
30
25

i/mA cm-2
30 20
15
10

-2
7

i/mA cm
20 5
0 6
0.0 0.5 1.0 1.5 2.0 2.5
Curea /M 5
10
4
23
0 1

0.2 0.3 0.4 0.5 0.6 0.7

E/V (Hg/HgO)

(b) 0
7

-1 100
Decrease in Qcathode / %

1
-2
i/mA cm

80

-2 60

40
0.0 0.5 1.0 1.5 2.0
Curea /M
-3
0.1 0.2 0.3 0.4 0.5
E/V (Hg/HgO)

Fig. 6. Effect of urea concentration on electrocatalytic oxidation of urea on nickel electrode in 1 M KOH solution at 10 mV s−1 : (a) cyclic voltammogram at various concentra-
tions of urea such (1) 0.5 M, (2) 0.4 M, (3) 0.3 M, (4) 0.2 M, (5) 0.1 M, (6) 0.05 M, and (7) 0 M. Inset shows the change in anodic peak current density at various concentrations
of urea. (b) The effect of urea concentration on cathodic peak in 1 M KOH. Inset shows the decrease in cathodic charge at various concentrations of urea (data taken from (a)).
The oxidation peak for urea increased with the addition of urea which is due to the high availability of urea molecules for oxidation (a). The inset shows that the anodic peak
current density increases linearly up to 0.2 M, after which it reaches constant values due to the increased coverage of the catalyst surface with urea and its intermediates,
which in turn decreases the oxidation rate of urea due to the local deprivation of OH− species. The results in (b) reveal the decrease in cathodic charge with increasing urea
concentrations indicating the decreased availability of NiOOH for reduction in the cathodic region.

shows a decrease in the cathodic charge in the presence of urea surface due to the occurrence of two parallel oxidation reactions in
at all the KOH concentrations tested. The decrease is linear up to the presence of urea in the KOH solution.
1 M KOH exhibiting a diffusion limitation of the OH− species to the
catalyst surface, after which it deviates from linearity. This devia- 3.1.4. Effect of urea concentration
tion at high concentrations of KOH (1–5 M) in the system may be Fig. 6 shows the effect of urea concentration on the electro-
mainly due to the complete coverage of the catalyst surface with catalytic oxidation of urea at a nickel electrode in 1 M KOH at a
OH− ions thus blocking the access of urea molecules for further oxi- scan rate of 10 mV s−1 . The peak ascribing to the oxidation of urea
dation. The percentage decrease in cathodic charge is high (50%) at increased with the addition of urea (Fig. 6a), whereas the cor-
low concentrations of KOH and low (39%) at high concentrations of responding cathodic charge decreased with the addition of urea
KOH. The diminution in cathodic charge in the presence of urea may (Fig. 6b). The increase in the anodic peak current density with
be related to the chemical reduction of NiOOH to Ni(OH)2 while oxi- increasing the concentration of urea is due to the high availability
dizing urea, as reported by other authors [22,45,46]. As discussed of urea molecules for the oxidation. As stated in Section 3.1.1, there
above, the decrease in the cathodic current density can also be asso- is a decrease in the Ni2+ wave upon addition of increasing amounts
ciated with the formation of less amount of NiOOH on the electrode of urea thus suggesting a possible EC mechanism. Also, the anodic
V. Vedharathinam, G.G. Botte / Electrochimica Acta 81 (2012) 292–300 297

-1 -1.5
440 mV
445 mV
450 mV
-2 2 M KOH -2.0
3 M KOH 455 mV

log/(i, mA cm )
-2
log/(i, A cm )
-2

4 M KOH
-3 -2.5

1 M KOH
-4 -3.0

-5 -3.5
0.25 0.30 0.35 0.40 0.45 0.50 0.55 -0.2 0.0 0.2 0.4 0.6 0.8

E/V (Hg/HgO) log/(CKOH, M)


Fig. 7. Tafel plots at various KOH concentrations in 0.33 M urea solution at a scan Fig. 8. Double logarithmic plot of current density as a function of KOH concentration
rate of 1 mV s−1 . The plot shows the enhanced onset potential with increased OH− at constant electrode potentials: () 440 mV; (䊉) 445 mV; () 450 mV; () 455 mV.
concentration. Experimental conditions as in Fig. 7. The plot is linear at all tested potentials with a
slope of 2, suggesting a reaction order of 2 with respect to OH− ions.

peak current density increases linearly up to 0.2 M urea. It could be


hypothesized that the increase is due to a diffusion-controlled pro- tested potentials with a slope of 2, suggesting a reaction order of 2
cess at low concentrations of urea. At concentrations of urea higher with respect to OH− ions.
than 0.2 M, the change in the anodic current density decreases and The reaction order for urea oxidation in 1 M KOH is obtained
becomes independent of the urea concentration that may be due under quasi-steady state conditions by varying the concentration
to a kinetics limitation control process. This phenomenon explains of urea (from 0.2 M to 2 M) at 1 mV s−1 and 25 ◦ C. Fig. 9 shows the
that the surface coverage of urea molecules on nickel becomes high Tafel plots using the above mentioned conditions. The average Tafel
at high concentrations, which in turn decreases the oxidation rate slope value obtained is 31 mV per current decade. A logarithmic plot
of urea due to the local deprivation of OH− species. Also, the peak of the current density vs. urea concentration at constant electrode
potentials corresponding to the oxidation of urea increases linearly potential is shown in Fig. 10. From the slope of the double logarith-
with increasing urea concentration over the range studied. Increas- mic plot, the reaction order is found to be 0.3 with respect to urea
ing urea concentration tends to increase the surface coverage of and is independent of the potential range considered.
urea and its intermediates, thus leading to the high anodic cur-
rent density by oxidizing the high amount of adsorbed urea and its 3.1.6. Rotating disk electrode (RDE) voltammetry
intermediates on the catalyst surface. In turn the adsorbed OH− ion To eliminate the possible interferences from mass transport
coverage on the catalyst surface will decrease due to the high sur- issues on the measured currents, the CV experiments were repeated
face coverage of urea and its intermediates. Hence, the oxidation of using a rotating disk electrode arrangement. Polarization curves
Ni2+ species to Ni3+ is achieved only by shifting the peak potentials recorded at different rotation speeds for the oxidation of urea on a
to more positive values. The inset in Fig. 6b shows the varia- nickel electrode in 5 M KOH solution containing 0.33 M urea are
tion in cathodic charge associated with the reduction of NiOOH to shown in Fig. 11. Though the shape of the polarization curves
Ni(OH)2 while varying the concentration of urea. The result shows
a decrease in cathodic charge with increasing urea concentrations.
-2.0
The excess availability of urea molecules at higher concentration
tends to reduce more NiOOH to ␤-Ni(OH)2 , as per reaction (5), in
the forward scan and the initial stages of the reverse scan. Hence, 1M
-2.4
the amount of NiOOH available for reduction in the cathodic region
0.8 M
log/(i, mA cm )

decreases, which is reflected in the decreased cathodic charge.


-2

-2.8 0.6 M
3.1.5. Tafel analysis
The LSV experiments were performed by varying OH− or urea 0.4 M
concentration in 0.33 M urea and 1 M KOH, respectively. Fig. 7 -3.2
0.2 M
shows the Tafel plot obtained under quasi-steady state conditions
for the urea oxidation reaction at various concentrations of KOH
-3.6
and 0.33 M urea at 1 mV s−1 and 25 ◦ C. The plot shows that the
onset potential increases with increasing OH− concentration. This
is consistent with the above obtained result in which the high OH− -4.0
concentration enhances the onset potential and the oxidation rate 0.40 0.42 0.44 0.46 0.48 0.50
of urea (Figs. 4 and 5). There is no significant change in the Tafel
slope with varying the OH− concentration and the average value E/V (Hg/HgO)
obtained is 28 mV per current decade. Fig. 8 shows the logarithmic
Fig. 9. Tafel plot at various concentrations of urea in 1 M KOH solution at a poten-
plot of the current density as a function of the OH− concentration tial scan rate of 1 mV s−1 . The plot does not show any enhancement shows any
at different constant electrode potentials. The plot is linear at all enhancement in the onset potential with increased concentration of urea.
298 V. Vedharathinam, G.G. Botte / Electrochimica Acta 81 (2012) 292–300

-2.5 150
440 mV
445 mV
450 mV 120
-3.0
log/(i, mA cm )
-2

i/mA cm-2
90
-3.5

60

-4.0
30

-4.5
0
-0.8 -0.6 -0.4 -0.2 0.0 0.2
0 10 20 30 40 50
log/(Curea, M)
f1/2/rpm1/2
Fig. 10. Double logarithmic plot of current density as a function of urea concen-
Fig. 12. Levich plot for electrocatalytic oxidation of urea on nickel electrode in 5 M
tration at constant electrode potentials: () 440 mV; (䊉) 445 mV; () 450 mV.
KOH containing 0.33 M urea (data taken from Fig. 11). The plot shows linear depen-
Experimental conditions as in Fig. 9. The slope of the double logarithmic plot at
dence at lower rotation speeds suggesting that the electrocatalytic oxidation of urea
constant electrode potential shows a reaction order of 0.3 with respect to urea.
on nickel electrode is controlled by diffusion. The deviation from linearity at high
rotation speeds implies that the reaction is limited by kinetics.

at different rotation speeds is similar, the recorded currents are


The polarization curves were analyzed according to the charac-
dependent on the rotation rate. The inset shows linearity until
teristics of an irreversible reaction controlled by diffusion, where
500 rpm indicating that the urea oxidation reaction is controlled
the current I (A) is given by the Koutecky–Levich equation [42]:
by mass transport limitations after which is mainly controlled by
kinetics limitations. Also the polarization curve at 2000 rpm (not 1 1 1
= + (8)
limited by mass transport) shows that the current density reaches I Ik ID
a maximum at ∼0.5 V vs. Hg/HgO during forward scan and then Ik (A) and ID (A) represent the kinetic and diffusion controlled cur-
decreases without the definition of a ‘plateau’ that is typical for rents, respectively, with
rotating disk voltammetry. This behavior can be due to the surface
2/3
blockage of the NiOOH catalyst during the urea oxidation reaction. ID = (0.62nFADo C v−1/6 )ω1/2 (9)
At higher potentials, the current density in the reverse scan is some-
where ‘n’ is the number of electrons transferred per reactant
what smaller than those on the forward scan, which may suggest
molecule, ‘F’ is the Faradaic constant (C mol−1 ), ‘A’ is the electrode
that the previously blocked active sites for urea oxidation were not
area (cm2 ), ‘Do ’ is the diffusion co-efficient (cm2 s−1 ), ‘C’ is the reac-
completely regenerated.
tant concentration in bulk solution (mol L−1 ), ‘v’ is the kinematic
viscosity of the solution (cm2 s−1 ), and ‘ω’ is the electrode rota-
tion rates (rad s−1 ). The constant used in the equation corresponds
300
150 with rad−1/2 . For all calculations, the angular velocity is considered
as ω = (2␲/60)f, where f is the frequency of the electrode rotation
250 (rpm).
i/mA cm-2

100
Fig. 12 shows the Levich plot for the oxidation of 0.33 M urea
200 in 5 M KOH at various rotation speeds. The results show linear
50 dependence at lower rotation speeds suggesting that the electro-
i/mA cm-2

5 catalytic oxidation of urea on a nickel electrode is controlled by


150 diffusion. Interestingly, the plot deviates from linearity at high rota-
0
0 500 1000 1500 2000 4 tion speeds implying that the reaction is limited by kinetics [42].
ω/ rpm 3
100 2 The electrocatalysis cannot keep up with the increased flux of urea
to the electrode surface and hence the anodic peak current density
reaches a plateau region at high rotation speeds.
50
1 According to Eqs. (8) and (9), the Koutecky–Levich (i−1 vs. ω−1/2 )
plot should be linear with a ‘zero’ intercept confirming the effects
0 of diffusion on the recorded currents at a given electrode potential,
as seen in Fig. 13. The intercepts recorded from the plot at different
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
potentials are always greater than zero indicating that the elec-
E/V (Hg/HgO) trocatalytic oxidation of urea in alkaline media is a mixed control
mechanism, where the currents are controlled by both diffusion
Fig. 11. Voltammograms of nickel electrode in 5 M KOH solution containing 0.33 M and electrode kinetics.
urea using a rotating disk electrode at various rotation speeds: 100, 300, 500, 1000,
and 1500 rpm. Inset shows the dependence of anodic current density on different 3.1.7. Mechanistic consideration
rotation speeds for urea oxidation on nickel electrode. The voltammograms reveal
As per the available literature, oxidation of organic molecules
that the anodic current density reaches a maximum during forward scan after which
it decreases without attaining the characteristic “plateau” region suggesting surface on nickel electrodes can be based on two mechanisms: (1) a direct
blockage of Ni catalyst. electron transfer from the organic compounds (org) to oxide films
V. Vedharathinam, G.G. Botte / Electrochimica Acta 81 (2012) 292–300 299

0.24 Net anodic reaction:

0.47 V CO(NH2 )2(aq) + 6OH− → N2(g) + 5H2 O(l) + CO2(g) + 6e− EC
0.20
The observations also suggest the possibility of a direct electro-
oxidation mechanism where the electrochemical oxidation of urea
0.16
(Eq. (2)) occurs on the surface of NiOOH. The decrease in cathodic
i-1/mA-1 cm-2

charge upon addition of urea may as well be attributed to the for-


0.12 mation of less amount of NiOOH on the electrode surface due to the
0.48 V occurrence of the two parallel oxidation reactions (Eqs. (1) and (2))
in the presence of urea in KOH solution when compared with KOH
0.08
solution alone (Eq. (1)). Hence the possibility of a direct oxidation
mechanism of urea on a nickel electrode cannot be ruled out. Fur-
0.49 V
0.04 ther analysis involving in situ surface characterization techniques is
0.50 V recommended to discriminate between the two different hypothe-
sized mechanisms: indirect oxidation and direct electro-oxidation.
0.00
0.04 0.08 0.12 0.16 0.20 0.24 0.28
4. Conclusion
ω-1/2/s-1/2
The electrocatalytic oxidation mechanism and kinetics of urea
Fig. 13. Koutecky–Levich plot for the oxidation currents measured at different elec-
on a nickel electrode in alkaline medium was investigated. The
trode potentials (0.41, 0.42, 0.43, and 0.44 V) in 5 M KOH containing 0.33 M urea (data
taken from Fig. 11). The intercepts recorded from the plot at different potentials are
electro-catalytic oxidation of urea occurs after the formation of
always greater than zero indicating that the electrocatalytic oxidation of urea in some amount of Ni3+ on the electrode surface. The oxidation also
alkaline media is a mixed control mechanism, where the currents are controlled by continues during the initial stages of reverse scan suggesting the
both diffusion and electrode kinetics. regeneration of active sites for further adsorption and oxidation
of urea or its intermediates. At low concentrations of urea, the
[47] (Eq. (10)) and (2) an indirect oxidation (catalyst regeneration- urea oxidation reaction is mainly controlled by diffusion whereas
EC mechanism) of the organic compounds through Ni3+ owing to at high concentrations the process is mainly controlled by kinetics.
the formation of Ni2+ (Eq. (5)) [35,38]: From the obtained data it is unlikely to propose a concrete reac-
tion mechanism for the electrochemical oxidation of urea on nickel
NiOOH + org → NiOOH(org)ads → NiOOH + products + H+ + e− electrodes. Few possibilities cannot be excluded. The decrease in
(10) cathodic charge when urea is present in KOH solution suggests
a possible indirect oxidation mechanism. But, this decrease can
Since there is no available literature about the electrocatalytic oxi- also be attributed to the usage of applied potential between two
dation mechanism of urea on nickel electrodes, the results will be oxidation reactions as (a) oxidation of Ni(OH)2 to NiOOH, and (b)
compared with the literature for oxidation of organic molecules oxidation of urea to CO2 and N2 in the presence of urea in KOH
on nickel based electrodes. Fleischm et al. [48,49] suggested that solution. The effect of scan rate shows the presence of diffusion (Ipa
NiOOH acts as an electrocatalyst for the oxidation of organic vs. 1/2 ) and kinetically (Epa vs. ) controlled mechanisms for the
molecules and it involves charging of the oxide layer followed by electro-oxidation reaction of urea. The results from RDE voltam-
a chemical reaction between NiOOH and the organic molecule. metry conclude that the reaction is controlled by diffusion until
This conclusion was based on the observation that the organic 500 rpm after which the reaction is controlled predominantly by
molecule oxidized at a potential that coincided exactly with the kinetics. The non-linearity at higher rotation speeds in the Levich
potential where NiOOH was formed and the cathodic peak corre- plot implies that the reaction is limited by electrode kinetics and
sponding to nickel hydroxide disappeared completely during the not by mass-transport. The presence of non-zero intercepts at
reverse scan. Vertes et al. [50] questioned the role of NiOOH in the different electrode potentials in Koutecky–Levich plot proves the
potential dependent oxidation of alcohols at oxide covered nickel presence of a mixed control mechanism, i.e., the urea oxidation
electrodes due to the presence of a separate redox peak corre- currents are controlled by both diffusion and electrode kinetics lim-
sponding to Ni(OH)2 /NiOOH followed by the appearance of a new itations. The kinetic parameters such as charge transfer coefficient
peak corresponding to alcohol oxidation at more positive poten- (˛) and the diffusion coefficient (Durea ) were calculated as 0.78 and
tials. Robertson [51] and Taraszewska et al. [52] addressed the same 0.85 × 10−5 cm2 s−1 , respectively. The Tafel studies show a reaction
issue. El-Shafei [53–55] suggested that the oxidation of methanol order of 2 and 0.3 with respect to OH− and urea, respectively.
on Ni(OH)2 /glassy carbon modified electrode in alkaline medium
occurred via Ni3+ species. Danaee et al. [54,55] reported that at Acknowledgements
high current density a part of the anodic current corresponding to
oxidation of glucose on NiOOH is due to indirect electro-oxidation The authors would like to thank the financial support of the Cen-
mechanism and the other part is due to a direct electro-oxidation ter for Electrochemical Engineering Research at Ohio University,
mechanism. and the Department of Defense through the U.S. Army Construc-
On the basis of the above reported results and the available tion Engineering Research Laboratory (W9132T-09-1-0001). The
literature, it is possible to hypothesize that electrochemical oxida- content of the information does not reflect the position or the pol-
tion of urea undergoes indirect oxidation mechanism (EC ) which icy of the U.S. government. The authors also thank Dr. Madhivanan
involves the reduction of NiOOH to Ni(OH)2 and electrochemical Muthuvel for his helpful discussion during the project.
regeneration of Ni(OH)2 to NiOOH, as summarized below:
References
6Ni(OH)2(s) + 6OH−  6NiOOH(s) + 6H2 O(l) + 6e− E
[1] B.K. Boggs, R.L. King, G.G. Botte, Chemical Communications (2009) 4859.
[2] G.G. Botte, U.S. Patent Application No. 0095636 A1 (2009).
6NiOOH(s) + CO(NH2 )2(aq) +H2 O(l) → 6Ni(OH)2(s) + N2(g) + CO2(g) C [3] D. Das, T.N. Veziroglu, International Journal of Hydrogen Energy 26 (2001) 13.
300 V. Vedharathinam, G.G. Botte / Electrochimica Acta 81 (2012) 292–300

[4] R.L. King, G.G. Botte, Journal of Power Sources 196 (2011) 2773. [32] J. Chen, D.H. Bradhurst, S.X. Dou, H.K. Liu, Journal of the Electrochemical Society
[5] F. Vitse, M. Cooper, G.G. Botte, Journal of Power Sources 142 (2005) 18. 146 (1999) 3606.
[6] C.C. Jara, S. Di Giulio, D. Fino, P. Spinelli, Journal of Applied Electrochemistry 38 [33] A.J. Motheo, S.A.S. Machado, F.J.B. Rabelo, J.R. Santos, Journal of the Brazilian
(2008) 915. Chemical Society 5 (1994) 161.
[7] J.Y. Jiang, M. Chang, P. Pan, Environmental Science and Technology 42 (2008) [34] R.M.A. Tehrani, S. Ab Ghani, Fuel Cells 9 (2009) 579.
3059. [35] C.-X. Cai, H.-X. Ju, H.-Y. Chen, Analytica Chimica Acta 310 (1995) 145.
[8] B.A. Till, L.J. Weathers, P.J.J. Alvarez, Environmental Science and Technology 32 [36] H.R. Zare, N. Nasirizadeh, S.M. Golabi, M. Namazian, M. Mazloum-Ardakani, D.
(1998) 634. Nematollahi, Sensors and Actuators B 114 (2006) 610.
[9] H.Q. Yu, Z.H. Zhu, W.R. Hu, H.S. Zhang, International Journal of Hydrogen Energy [37] C.P. Andrieux, J.M. Saveant, K.B. Su, Journal of Physical Chemistry 90 (1986)
27 (2002) 1359. 3815.
[10] H.S. Zhang, M.A. Bruns, B.E. Logan, Water Research 40 (2006) 728. [38] D. Nematollahi, M. Rafiee, L. Fotouhi, Journal of the Iranian Chemical Society 6
[11] L. Zhou, Y.F. Cheng, International Journal of Hydrogen Energy 33 (2008) 5897. (2009) 448.
[12] R.L. King, MSc Thesis, 2010. [39] M.-Z. Czae, J. Wang, Talanta 50 (1999) 921.
[13] R. Lan, S.W. Tao, J.T.S. Irvine, Energy & Environmental Science 3 (2010) 438. [40] I. Danaee, M. Jafarian, F. Forouzandeh, F. Gobal, M.G. Mahjani, International
[14] W. Simka, J. Piotrowski, G. Nawrat, Electrochimica Acta 52 (2007) 5696. Journal of Hydrogen Energy 33 (2008) 4367.
[15] A.P.S. Terblanche, Water SA 17 (1991) 77. [41] J.J. Zhang, Y.H. Tse, W.J. Pietro, A.B.P. Lever, Journal of Electroanalytical Chem-
[16] I. Danaee, M. Jafarian, F. Forouzandeh, F. Gobal, M.G. Mahjani, International istry 406 (1996) 203.
Journal of Hydrogen Energy 34 (2009) 859. [42] L.M.F. Dantas, A.P. dos Reis, S. Tanaka, J.H. Zagal, Y.Y. Chen, A.A. Tanaka, Journal
[17] M. Vidotti, M.R. Silva, R.P. Salvador, S.I.C. de Torresi, L.H. Dall’Antonia, Elec- of the Brazilian Chemical Society 19 (2008) 720.
trochimica Acta 53 (2008) 4030. [43] V. Gau, S.-C. Ma, H. Wang, J. Tsukuda, J. Kibler, D.A. Haake, Methods 37 (2005)
[18] R.M.A. Hameed, K.M. El-Khatib, International Journal of Hydrogen Energy 35 73.
(2010) 2517. [44] F. Pariente, E. Lorenzo, F. Tobalina, H.D. Abruna, Analytical Chemistry 67 (1995)
[19] G.P. Jin, Y.F. Ding, P.P. Zheng, Journal of Power Sources 166 (2007) 80. 3936.
[20] Q.F. Yi, W. Huang, W.Q. Yu, L. Li, X.P. Liu, Electroanalysis 20 (2008) 2016. [45] H. Heli, M. Jafarian, M.G. Mahjani, F. Gobal, Electrochimica Acta 49 (2004)
[21] Q.F. Yi, J.J. Zhang, W. Huang, X.P. Liu, Catalysis Communications 8 (2007) 1017. 4999.
[22] S. Majdi, A. Jabbari, H. Heli, Journal of Solid State Electrochemistry 11 (2007) [46] M. Jafarian, F. Forouzandeh, I. Danaee, F. Gobal, M.G. Mahjani, Journal of Solid
601. State Electrochemistry 13 (2009) 1171.
[23] E. Nicolau, I. Gonzalez-Gonzalez, M. Flynn, K. Griebenow, C.R. Cabrera, Advances [47] A. Kapalka, A. Cally, S. Neodo, C. Comninellis, M. Wachter, K.M. Udert, Electro-
in Space Research 44 (2009) 965. chemistry Communications 12 (2010) 18.
[24] A.M. Barrios, S.J. Lippard, Journal of the American Chemical Society 122 (2000) [48] M. Fleischm, K. Korinek, D. Pletcher, Journal of Electroanalytical Chemistry 31
9172. (1971) 39.
[25] D.A. Daramola, D. Singh, G.G. Botte, Journal of Physical Chemistry A 114 (2010) [49] M. Fleischm, D. Pletcher, K. Korinek, Journal of the Chemical Society, Perkin
11513. Transactions 2 (1972) 1396.
[26] F. Musiani, E. Arnofi, R. Casadio, S. Ciurli, Journal of Biological Inorganic Chem- [50] G. Vertes, G. Horanyi, Journal of Electroanalytical Chemistry 52 (1974)
istry 6 (2001) 300. 47.
[27] J. Desilvestro, D.A. Corrigan, M.J. Weaver, Journal of the Electrochemical Society [51] P.M. Robertson, Journal of Electroanalytical Chemistry 111 (1980) 97.
135 (1988) 885. [52] J. Taraszewska, G. Roslonek, Journal of Electroanalytical Chemistry 364 (1994)
[28] C.J. Zhang, S.M. Park, Journal of the Electrochemical Society 136 (1989) 3333. 209.
[29] M.A. Hopper, J.L. Ord, Journal of the Electrochemical Society 120 (1973) 183. [53] A.A. El-Shafei, Journal of Electroanalytical Chemistry 471 (1999) 89.
[30] A. Van der Ven, D. Morgan, Y.S. Meng, G. Cederc, Journal of the Electrochemical [54] I. Danaee, M. Jafarian, F. Forouzandeh, F. Gobal, M.G. Mahjani, Journal of Physical
Society 153 (2006) A210. Chemistry B 112 (2008) 15933.
[31] A.N. Golikand, M.G. Maragheh, L. Irannejad, M. Asgari, Russian Journal of Elec- [55] I. Danaee, M. Jafarian, F. Forouzandeh, F. Gobal, M.G. Mahjani, Electrochimica
trochemistry 42 (2006) 167. Acta 53 (2008) 6602.

You might also like