You are on page 1of 7

Journal of The Electrochemical Society, 166 (10) B859-B865 (2019) B859

0013-4651/2019/166(10)/B859/7/$38.00 © The Electrochemical Society

Free-Standing Copper Oxide Films with In Situ Tailorable


Nanostructures toward Detection of Hydrogen Peroxide
1,2,z
Peng Gao and Yuxuan Gong2,3,z
1 College of Materials Science and Engineering, Hunan University, Changsha 410082, China
2 Kazuo Inamori School of Engineering, New York State College of Ceramics at Alfred University, Alfred, New York
14802, USA
3 Owens Corning Science and Technology Center, Owens Corning, Granville, Ohio 43023, USA

In this study, in-situ synthesized Cu(OH)2 nanostructures on copper foil substrates were realized via anodization method. With
the precise manipulation of deposition conditions, qualitative modulation of the dimensionality of Cu(OH)2 nanostructures were
achieved. The influences of NaOH concentration, current density, reaction temperature and time on the morphologies of Cu(OH)2
nanostructures were investigated using complementary techniques. Furthermore, the as-prepared Cu(OH)2 nanostructures can be
successfully transformed to CuO with unnoticeable changes in morphology after thermal treatment in nitrogen atmosphere; therefore,
leading to a freestanding electrode for integrated-sensors. The resultant morphology and crystallographic identification of the as-
prepared samples were studied using scanning electron microscopy (SEM) and X-ray diffraction (XRD). It was found that the
pine-needle-like CuO modified electrode exhibited excellent electrochemical catalytic performance toward H2 O2 reduction. Highly
sensitive and fast-and-linear responses were achieved with its large specific surface area and efficient charge transfer properties.
Hence, we report the freestanding CuO films with facile tailorable nanostructures as a promising candidate for efficient, stable and
highly sensitive non-enzymatic amperometric detection of H2 O2 .
© 2019 The Electrochemical Society. [DOI: 10.1149/2.0131912jes]

Manuscript submitted May 6, 2019; revised manuscript received June 18, 2019. Published July 1, 2019.

The rapid, accurate and reliable detection of hydrogen peroxide nanomaterials in the precursor solution. Second, nanomaterials are
(H2 O2 ), a common strong oxidant and essential intermediate in many usually distributed randomly on the electrode surface after fabrica-
biomedical reactions,1 is of great significance.2–4 Among reported tion processes, therefore leads to a spatially non-uniform morphology
techniques for the detection of H2 O2 ,5–8 the electrochemical method and unpredictable catalytic performance. Third, the nanomaterials can
appears to be most promising due to its high sensitivity, fast response, easily detach from the electrode surface upon contact with the elec-
facile fabrication and operation, and low cost.9,10 In the past few trolyte, resulting in poor electrode stability. Therefore, several other
decades, enzyme-based electrochemical sensors such as horseradish approaches have been proposed for preparing stable and spatially ho-
peroxidase (HRP),11 glucose oxidase (GOx ),12 glucose dehydroge- mogeneous electrodes, such as electrodeposition24 and self-assembly
nases (GDHs),13 fructose dehydrogenase (FDH),14 d-amino acid oxi- methods.25 Among these proposed techniques, direct in-situ growth
dase (DAAO),15 etc. have attracted quite some interests due to good se- of Cu(OH)2 nanostructures on Cu foil substrate (achieved by anodiza-
lectivity. However, enzyme-based sensors as such usually suffer from tion and subsequent transformation to CuO through heat-treatment in
stability issues due to the fact that they are prone to be damaged ei- nitrogen) provides an alternative approach for obtaining catalytic elec-
ther thermally or chemically during fabrication and testing.16 Thus, trodes, given the fact that it provides a facile and scalable approach
attempts in resolving such issues have been made with a focus on to produce nanostructures with large specific surface areas and high
developing enzyme-free electrochemical sensors. electron charge transfer efficiency. However, the in-situ formation of
Copper oxide (CuO), a typical p-type semiconductor with a narrow Cu(OH)2 /CuO nanostructures on Cu foil substrate via anodization ap-
bandgap of 1.2 eV, has emerged to be one of the promising candidates proach, and their application in non-enzymatic amperometric deter-
as an active electrode material for enzyme-free electrochemical sen- mination of H2 O2 , have yet to be reported systematically.
sors. Its candidacy can be attributed to its excellent electrochemical In this paper, enzyme-free H2 O2 electrochemical sensors based on
catalytic activity and good chemical stability.17 The manipulation of different CuO nanostructures were prepared using a facile and versatile
morphologies in nanoscale is considered as an effective method of anodization method. The effects of NaOH concentration, current den-
improving the sensing performance of CuO toward H2 O2 , i.e. mor- sity, reaction temperature and time on the morphology of Cu(OH)2
phologies in nanoscale possess high surface area that directly trans- nanostructures were systematically investigated. Furthermore, their
lates to available sites for surface absorption.18,19 Therefore, efforts CuO counterparts show no morphological change after heat-treatment
have been focused on preparing nanostructured CuO and its appli- in nitrogen atmosphere. Enhanced electrochemical catalytic perfor-
cation in enzyme-free amperometric detection of H2 O2 . Qian20 et mance for H2 O2 reduction was obtained on the pine-needle-like CuO
al. prepared an urchin-like core-shell CuO nanostructure that was modified electrode, given its large specific surface area and efficient
assembled by closely packed nanorods through hydrothermal reac- electron charge transfer. All experimental results demonstrated that
tions, and the as-prepared sensors exhibited a wide range of capability the freestanding pine-needle-like CuO electrode exhibited great ap-
toward the detection of H2 O2 with a linear response (concentration plication potential in electrochemical sensing devices.
from 10 μM to 5.55 mM). Whang21 et al. reported a flower-like CuO
nanostructure on Cu foil via oxidation of copper under hydrothermal
conditions. The as-prepared sensors showed good catalytic perfor-
mance in detecting H2 O2 under alkaline conditions with a sensitivity of Materials and Methods
88.4 μA/mM cm2 . Wang22 et al. synthesized a hollow-structured cop- Chemicals and reagents.—High-purity copper foil (Cu, 99.9%),
per oxide by electrospinning, which demonstrated high sensitivity NaH2 PO4 •H2 O, Na2 HPO4 , sodium chloride (NaCl), ascorbic acid
(1746.50 μA mM−1 cm−2 ) and low detection limit (0.022 μM) for (AA), uric acid (UA), and dopamine (DA) were purchased from Alfa
the detection of H2 O2 . Aesar. Potassium chloride (KCl) was obtained from Sigma-Aldrich.
Conventionally, nanostructured CuO powders are bonded onto the Potassium ferricyanide (K3 Fe(CN)6 ), hydrogen peroxide solutions
electrode surface via film-casting or drop-coating methods, which have (H2 O2 , 30%) and sodium hydroxide (NaOH) were purchased from
shown certain disadvantages.23 First, it is difficult to evenly disperse Fisher Scientific. All reagents were of analytical grade, and being
used without further purification. Phosphate buffer solutions (PBS)
z
E-mail: gaop@hnu.edu.cn; pg3@alfred.edu; gongyauc@gmail.com; were prepared by mixing 0.1 M Na2 HPO4 and 0.1 M NaH2 PO4 •H2 O
yuxuan.gong@owenscorning.com solutions, and adjusting the pH value with 0.1 M H3 PO4 or 0.1 M

Downloaded on 2019-07-24 to IP 132.174.251.90 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
B860 Journal of The Electrochemical Society, 166 (10) B859-B865 (2019)

NaOH. H2 O2 solutions were diluted immediately before the electro-


chemical measurements.

In-situ preparation of various CuO nanostructures modified


electrodes.—The in-situ synthesis of freestanding CuO nanostructures
was achieved using an anodization method. In a typical procedure, a
piece of copper foil (10 mm × 10 mm × 0.127 mm) was washed with
distilled water and ethanol to remove surface impurities before use.
The Cu(OH)2 nanostructures were grown onto the Cu foil via direct
anodization of copper in a 2 M NaOH solution under a constant current
density of 4 mA/cm2 for 750 s at room temperature in a three electrode
system. An Ag/AgCl and a platinum wire were used as the reference
and auxiliary electrodes respectively. After reaction, a faint-blue film
was formed on Cu foil surface, which was washed with distilled water
and dried in air prior to any characterization. The resultant Cu(OH)2
nanostructures as a function of NaOH concentration, current density,
reaction temperature and time were studied. In order to obtain CuO
nanostructures, the Cu(OH)2 samples were heated at 200°C for 3 h in
nitrogen atmosphere. After heat-treatment, a dark film of CuO was
formed on the Cu foil.

Apparatus and measurements.—All the electrochemical experi-


ments were carried out using a CHI 650E electrochemical analyzer Figure 1. SEM images of a copper surface after anodizing at a constant current
(CHI, USA) with a three-electrode cell (same set-up, reference and density of 4 mA/cm2 in different concentrations of NaOH at room temperature
auxiliary electrodes as used for anodization process). The freestanding (25°C): (A) 1 M NaOH; (B) 2 M NaOH; (C) 3 M NaOH; (D) 4 M NaOH.
CuO nanostructures modified Cu electrodes were used as the working
electrode. 0.1 M KCl was used as supporting electrolyte in a 0.1 M
phosphate buffer solution (pH = 7.4). Electrochemical impedance samples’ morphologies and crystal structures. Generally speaking, the
spectroscopy (EIS) measurements were performed in 0.1 M KCl so- anodization process includes electro-oxidation of Cu to Cu2+ , which is
lutions within the frequency range of 0.1 Hz to 100 kHz using 5.0 mM then combined with OH− in the electrolyte solution to form Cu(OH)2
K3 Fe(CN)6 /K4 Fe(CN)6 (1:1) mixture as the electroactive probe. The (Eqs. 1–2).26 However, the Cu(OH)2 is a metastable phase, which
electrolytic solutions were purged with purified nitrogen for at least easily transforms to more stable CuO (Eq. 3).27 Therefore, the dehy-
30 min before electrochemical measurements. All the experiments dration of Cu(OH)2 to CuO is competitive to the growth of Cu(OH)2
were carried out at room temperature. nanocrystals during the nucleation process.26 At low NaOH concen-
Crystal structures were characterized by powder X-ray diffraction tration (1 M), the amount of generated Cu(OH)2 was not sufficient for
(XRD) via a Bruker D2 diffractometer at the voltage of 30 kV and further growth. Thus, it tends to completely dehydrated to CuO and
current of 10 mA with Cu Kα radiation (λ = 0.15418 nm), in a 2θ forming aggregate morphology. 2 and 3 M NaOH concentration were
angular range from 10 to 75° at a scanning speed of 9°/min. Sur- enough for nucleation and crystal growth, and thereby leading to var-
face morphologies were observed by a scanning electron microscopy ious Cu(OH)2 nanostructures. With further increasing NaOH concen-
(SEM, FEI Quanta 200) operating at 20 kV. tration to 4 M, the dehydration rate of Cu(OH)2 to CuO accelerated,26
and the reactions shown in Eqs. 4–5 happened, leading to the recon-
structive transformation of Cu(OH)2 to CuO28 and results in the ag-
Results and Discussion gregated CuO particles with few sparsely grown Cu(OH)2 nanorods.
Formation and characterization of different Cu(OH)2 /CuO Therefore, the nanostructures of resultant films can be manipulated
nanostructures.—The effects of NaOH concentration on morphology via the control of NaOH concentration in precursors, and further ma-
and crystal structure of as-prepared samples are shown in Figure 1. A nipulate the catalytic H2 O2 sensing properties.
series of experiments were carried out by varying the NaOH concen-
tration from 1 to 4 M (a constant current density of 4 mA/cm2 was Cu → Cu2+ + 2e− [1]
used for all experiments). As shown in Figure 1A, a black film of −
Cu 2+
+ 2OH → Cu(OH )2 [2]
densely aggregated nanoparticles was formed when the concentration
of NaOH is 1 M. Such dense film was determined to be a monoclinic Cu(OH )2 → CuO + H2 O [3]
CuO (PDF Card No. 01-080-1916) using XRD (Fig. S1, Supplemen- −
Cu(OH )2 + 2OH → Cu(OH )2−
4 [4]
tary material). A pine-needle-like Cu(OH)2 nanorods with a relatively

4 → CuO + 2OH + H2 O
large aspect ratio was formed with 2 M NaOH used, as shown in Cu(OH )2− [5]
Figure 1B. Each nanorod possessed a near-identical diameter and
length, which were approximately 150 nm and 10 μm, respectively. In addition, the effect of anodization current density on the mor-
The pine-needle-like Cu(OH)2 nanorods tend to self-assemble into a phology of resultant samples was studied. The current density was
flower-like structure on copper foil substrates. Further increment of varied from 2 to 6 mA/cm2 in a 2 M NaOH solution at room tempera-
NaOH concentration to 3 M developed a randomly oriented Cu(OH)2 ture. As shown in Figure 2A, samples prepared under a current density
nanoneedle structure (see Figure 1C), which were about 300 to 320 nm of 2 mA/cm2 were almost compactly aggregated CuO nanoparticles.
in diameter at its root and 130 to 140 nm at its tip. The diffraction pat- This is also attributed to the fact that the growth of Cu(OH)2 com-
terns of 2 M and 3 M samples matched well with Cu(OH)2 phase (PDF peted with the dehydration of Cu(OH)2 to CuO during the nucleation
Card No. 00-003-0307) (see Figure S1 in Supplementary material). As process. When a current density of 2 mA/cm2 was applied, gener-
shown in Figure 1D, with the NaOH concentration increased to 4 M, ated Cu(OH)2 nuclei was not sufficient enough for further growth.
a film of compactly aggregated CuO particles with very few short Hence, the resultant crystalline phase tends to be CuO (due to its
Cu(OH)2 nanorods (3-4 μm in length) sparsely grown on the CuO high reaction activity).26 After increasing the current density to 3
surface was formed (CuO surface on the Cu substrate was identified or 4 mA/cm2 , pine-needle-like Cu(OH)2 nanorods with some self-
via XRD, shown in Figure S1 in Supplementary material). The above assembled flowers were formed. Flabellate-like Cu(OH)2 structures
results clearly indicate important role of NaOH concentrations on the with few large-aspect-ratio nanoneedles were formed under a current

Downloaded on 2019-07-24 to IP 132.174.251.90 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 166 (10) B859-B865 (2019) B861

Figure 2. SEM images of a copper surface after anodizing in a 2 M NaOH Figure 3. SEM images of samples anodized at a constant current density of
solution (25°C) under different current densities: (A) 2 mA/cm2 ; (B) 3 mA/cm2 ; 4 mA/cm2 in a 2 M NaOH solution with different reaction temperatures: (A)
(C) 5 mA/cm2 ; (D) 6 mA/cm2 . 40°C; (B) 60°C; (C) 80°C; Figure (D) shows their corresponding XRD patterns:
(a) 40°C; (b) 60°C; (c) 80°C.

density of 5 mA/cm2 , as shown in Figure 2C. The nanoneedles were


about 300 to 310 nm in diameter at their roots and 110 to 120 nm at their voltage/current density, electrolyte concentration, reaction tempera-
tips. When the current density was further increased to 6 mA/cm2 , sim- ture, etc.), therefore different start time of film formation were reported
ilar flabellate-like Cu(OH)2 structures were obtained (see Figure 2D). in several published research works.30–32 The flabellate-like structure
However, the resultant nanoneedles appeared to be much shorter than (w/ similar length and diameter) tended to disappear when the anodiza-
the ones in 5 mA/cm2 sample. The typical XRD patterns of as-prepared tion time was further increased (see Figure 4B) to 600 s. Pine-needle-
samples are shown in Figure S2 (Supplementary material). All diffrac- like nanorods and randomly-oriented nanoneedles of Cu(OH)2 were
tion peaks of the samples prepared under 3–6 mA/cm2 current densities formed in the anodization time range of 750 to 1000 s and 1500
can be indexed to the Cu(OH)2 phase (PDF Card No. 00-003-0307), to 2000 s, respectively (see Figures 4C to 4F). XRD patterns of as-
whereas the sample obtained by using 2 mA/cm2 current density cor- prepared samples are shown in Figure S3 (Supplementary material).
responding to monoclinic CuO phase (PDF Card No. 01-080-1916). All measured diffraction peaks were well indexed as Cu(OH)2 phase
Samples prepared under different anodization temperatures (40, (PDF Card No. 00-003-0307), indicating high purity of the as-prepared
60 and 80°C; in a 2 M NaOH solutions and a constant current den- samples.
sity of 4 mA/cm2 ) were further studied. The SEM images of these The representative pine-needle-like and honeycomb-like Cu(OH)2
three samples are shown in Figure 3. As shown in Figure 3A, a non- nanostructures were further heat-treated at 200°C for 3 h in a nitro-
porous structure with completely closed pores was obtained when the gen atmosphere (see Figure S4). Heat-treated CuO nanostructures
anodization was carried out at 40°C. When the electrolyte tempera- retained the morphology of their Cu(OH)2 precursors with a phase
ture was increased to 60°C, numerous open pores in irregular shapes transition from Cu(OH)2 to monoclinic CuO phase (PDF Card No. 01-
were formed (Figure 3B). The pores formed at 60°C were randomly 080-1916). Therefore, heat-treatment in nitrogen can be used to suc-
distributed, appeared to a honeycomb structure in its spatial distribu- cessfully transform Cu(OH)2 to CuO without destroying their unique
tion with a pore size around 100 to 200 nm. A further increment of morphologies.
the anodization temperature to 80°C developed a less porous structure Based on above experimental results, the tendency of forming var-
(see Figure 3C) with thicker pore-walls and smaller pore size than ious Cu(OH)2 /CuO nanostructures on copper foil substrates under dif-
that of Figure 3B. XRD analysis (see Figure 3D) indicated that all ferent anodization conditions (NaOH concentration, current density,
three samples matched well with the Cu(OH)2 phase (PDF Card No. anodization temperature and time) has been summarized in Figure 5.
00-003-0307). The formation of porous structures as such was reported When a NaOH concentration of 1 M was used, no films were formed
to a direct result of Cu(OH)2 growth being disturbed by generation and on the copper foil regardless of the current density and reaction time.
motion of oxygen during anodization.29 Densely compacted CuO nanoparticles (see Figure 5, lower-left region
Finally, samples prepared at a constant current density of 4 mA/cm2 under the diagonal) were formed when the ratio of the current to NaOH
(room temperature, 2 M NaOH solution) were anodized over time to concentration is ≤ 1. Various Cu(OH)2 nanostructures self-assembled
investigate the relationship between Cu(OH)2 nanostructures and an- by nanoneedles or nanorods were formed when a ratio as such is
odization time. Their corresponding SEM images (Figure 4) shows that ≥ 1. The aspect-ratio, morphology and diameter of each nanoneedle
the as-prepared samples exhibited well-organized and spatially homo- or nanorod were controlled by the anodization conditions (NaOH con-
geneous morphologies. A gradual change of the dimensions and yield centration, current density, and reaction time). Further increase of the
of Cu(OH)2 nanostructures with increasing anodization time was ob- anodization temperature resulted in a honeycomb-like porous struc-
served. At a short reaction time, no Cu(OH)2 /CuO nanostructures were ture of Cu(OH)2 . We believe such a diagram (Figure 5) will provide a
observed on the copper surface, since the nucleation and crystal growth preliminary criterion of choosing appropriate anodization conditions
processes are kinetically-controlled. When increased the anodization to prepare desired Cu(OH)2 /CuO nanostructures.
time to 450 s, flabellate-like Cu(OH)2 structures were formed on the
copper substrate in an uniform-but-loose fashion (Figure 4A). How- Determination of effective surface area and charge trans-
ever, it is worth noting that the start of the formation of Cu(OH)2 fer properties.—Aforementioned freestanding pine-needle-like,
nanostructures depends on the anodization conditions (such as honeycomb-like and aggregate CuO nanostructures, which represent

Downloaded on 2019-07-24 to IP 132.174.251.90 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
B862 Journal of The Electrochemical Society, 166 (10) B859-B865 (2019)

Figure 4. SEM images of the as-prepared Cu(OH)2 nanostructures obtained at a constant current density of 4 mA/cm2 in 2 M NaOH solution for different reaction
times at room temperature (25°C): (A) 450 s; (B) 600 s; (C) 750 s; (D) 1000 s; (E) 1500 s; (F) 2000 s. Insets are their corresponding magnified SEM images.

three typical morphologies obtained during anodization processes, more exposed surface absorption sites. The apparent smaller effective
were then directly used as the sensor electrodes and their electrochemi- surface area of aggregate CuO modified electrode demonstrates that
cal catalytic performances toward H2 O2 reduction were studied. First, the densely compacted film is less favorable for obtaining large specific
we estimated the electrochemical effective surface areas (A) of the surface area, which may diminish its electrochemical activity toward
three different electrodes using cyclic voltammetry, with ferricyanide H2 O2 detection.
as a redox probe. For each CuO modified electrode, a series of cyclic Electrochemical impedance spectroscopy (EIS) is an effective
voltammograms were recorded as a function of different scan rates method for probing the ion diffusion and electron charge transfer
(1 mM K3 Fe(CN)6 solution with 0.1 M KCl as supporting electrolyte). properties between the electrolyte and the electrode surface, and
As shown in Figures 6a–6c, the peak current changes linearly with the thereby was used to investigate the different CuO nanostructures mod-
square root of the scan rate regardless of the electrode morphology, ified electrodes.9 A 10 mL 0.1 M KCl solution containing 5.0 mM
indicating that the mass transfer process in the double layer region of K3 Fe(CN)6 /K4 Fe(CN)6 (1:1) was used as the supporting electrolyte.
the electrode surface is diffusion-controlled. The applied frequency was in the range of 0.1 Hz to 100 kHz under
The dependence of peak currents on scan rates can be described an open circuit potential of +0.22 V. As shown in Figure 6d, all three
by Randles-Sevcik equation under semi-infinite linear diffusion con- plots exhibited a straight line in the low frequency region and a single
ditions and room temperature (25°C):21 semicircle in the high frequency region, indicating a diffusion-limited
    step in the low frequency region and an electron-transfer-limited
Ip = 2.69 × 105 n3/2 AD0 1/2C0 υ 1/2 n = 1, D0 = 7.6 × 10−6 cm2 /s step in the high frequency region.9 The charge transfer resistance
[6] (Ret) estimated from the diameter of the semicircle is about 122 and
where Ip is the peak current, ν is the scan rate, n, D0 and C0 are 130  for pine-needle-like and honeycomb-like CuO modified elec-
constants. By using this equation, the electrochemical effective surface trodes respectively, which are smaller than that of aggregate CuO mod-
area can be calculated as 8.41, 7.77 and 3.06 cm2 corresponding to ified electrode (330 ). In addition, the steep slopes of pine-needle-like
pine-needle-like, honeycomb-like, and aggregate CuO respectively. and honeycomb-like CuO modified electrodes in the low frequency
The effective surface area of pine-needle-like CuO is slightly larger region suggested a lower resistance of electrolyte diffusion compar-
than that of honeycomb-like porous CuO, which is probably due to the ing to that of aggregate CuO electrodes. Thus, both the pine-needle-
like and honeycomb-like CuO modified electrodes could facilitate the
mass transport and electron charge transfer during electrochemical
processes33 when compared to aggregate CuO modified electrode.
Moreover, the Ret values of pine-needle-like and honeycomb-like CuO
were much smaller than the reported value of CuO nanomaterials mod-
ified glassy carbon electrodes using film-casting method.34–36 Hence,
the direct in-situ growth of CuO nanostructures on copper foil sub-
strate greatly enhanced the electron charge transfer efficiency during
electrochemical processes.

The amperometric detection of hydrogen peroxide.—The elec-


trochemical catalytic detections of H2 O2 were conducted for the
as-prepared three electrodes with i-t technique, in order to further
elucidate the correlation between different nanostructures and their
H2 O2 sensing performance. The catalytic mechanism typically in-
volves electrochemical reduction of Cu2+ to Cu+ , which is then chemi-
Figure 5. Schematic diagram showing the tendency of forming various cally reacted with H2 O2 and resulted in the conversion of H2 O2 to H2 O,
Cu(OH)2 /CuO nanostructures under different anodization conditions. accompanied by the regeneration of Cu2+ catalyst (see Figure S5).2

Downloaded on 2019-07-24 to IP 132.174.251.90 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 166 (10) B859-B865 (2019) B863

Figure 6. Peak current as a function of the square root of scan rates for (a) pine-needle-like CuO, (b) honeycomb-like CuO, and (c) aggregate CuO. Inset: CVs
of the three different CuO nanostructures modified electrodes in 1 mM ferricyanide solution containing 0.1 M KCl at different scan rates (20–300 mV/s). (d)
Electrochemical impedance spectroscopy of the three as-prepared CuO nanostructures (pine-needle-like, honeycomb-like and aggregate CuO corresponding to
curve a, b and c, respectively) in 10 mL 0.1 M KCl solution containing 5.0 mM K3 Fe(CN)6 /K4 Fe(CN)6 (1:1).

Figure 7 shows the typical amperometric response curves of all three The sensitivities were estimated, with the use of calibration curve, to
CuO nanostructures modified electrodes upon successive additions of be 740, 705 and 424 μA/mM for pine-needle-like, honeycomb-like,
H2 O2 into gently stirred PBS solutions. Clearly, the reduction currents and aggregate CuO modified electrodes. The analytical performances
increased immediately and reached their steady-state value (95% of of the as-prepared electrodes in this research were also compared with
the maximum) in less than 3 s upon injection of H2 O2 for all samples, others reported in the literature, and is shown in Table I. Apparently,
indicating fast and stable responses toward H2 O2 reduction. No am- our freestanding pine-needle-like CuO modified electrodes exhibited
perometric responses of dissolved oxygen in the buffer solution were much higher sensitivity than most drop-coated CuO nanostructures
observed for any of the three samples, which is probability due to the modified electrodes37–39 and several freestanding CuO electrodes,40–42
near neutral working solution conditions.35 Therefore, the responses given its unique morphology as well as large electrochemical effective
of H2 O2 can be measured without interference from dissolved oxygen. surface area. Such relationship between sensitivity and morphology

Table I. Comparison of various CuO nanomaterials modified electrodes for H2 O2 detection.

Modified materials Sensitivity (μA/mM) Linear range (mM) Detection limit (mM) References

Pine-needle-like CuO 740 0.25–5 0.12 This work


Honeycomb-like CuO 705 0.25–5 0.14 This work
Aggregate CuO 424 0.25–5 0.17 This work
Cubic Cu2 O 35.2 0.05–3.4 1.54 × 10−2 37
Cu2 O-rGO 15.5 0.1–9.8 0.079 38
CuO nanoleaves 573.1 0.019–5 1.0 × 10−2 39
CuO-GO 262.5 2.79 × 10−3 -2 0.69 × 10−3 40
Cu2 O/TiO2 /Ti 412.1 0.5–8 0.1 41
CuO Nanosheet 25.5 0.01–20 N/A 42

Downloaded on 2019-07-24 to IP 132.174.251.90 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
B864 Journal of The Electrochemical Society, 166 (10) B859-B865 (2019)

Figure 7. Amperometric responses and corresponding calibration curves between catalytic currents and H2 O2 concentrations of (a, d) pine-needle-like CuO, (b,
e) honeycomb-like CuO, and (c, f) aggregate CuO modified electrodes upon successive addition of H2 O2 into gently stirred 0.1 M PBS (pH = 7.4) containing
0.1 M KCl at an applied potential of −0.4 V.

agreed well with the observations obtained from EIS measurements as the sensor electrodes, and their electrochemical catalytic perfor-
and electrochemical effective surface areas. It is worth noting that mance toward H2 O2 reduction were investigated. It was found that the
the proposed synthetic approaches and as-prepared freestanding CuO pine-needle-like CuO modified electrode exhibited excellent catalytic
electrodes exhibit not only higher sensitivity, but also more advan- performance toward H2 O2 reduction, including high sensitivity (740
tages compared to most reported works. The in-situ grown of CuO μA/mM), fast response (< 3 s), good selectivity (toward the detec-
nanostructures on Cu foil substrate via anodization provides a facile tion of H2 O2 vs. common interfering species, such as AA, UA, DA,
and scalable approach to produce sensing electrodes, which enables and NaCl) and long-term stability. Performance as such was due to its
their practical applications (such as portable and wearable electron- large specific surface area, uniform microstructure, and efficient elec-
ics). The freestanding nature also promotes charge transfer efficiency tron charge transfer. Therefore, the advantages of facile and scalable
and therefore leading to better electrochemical catalytic performance. electrodes fabrication, as well as excellent electrochemical catalytic
More importantly, the quantitative assessment of different deposition performance made the self-standing pine-needle-like CuO modified
conditions leading to a practically useful diagram, which provides electrode one of the most promising candidates for the electrochemi-
a preliminary criterion of choosing appropriate anodization condi- cal detection of H2 O2 .
tions to get desired CuO nanostructures. We believe more researches
could be conducted in the future based on our work, such as preparing
heterostructures or incorporating structural defects, to further realize
better electrochemical catalytic performance.
Good selectivity and long-term stability of the sensing electrodes
are also important for their practical applications. Therefore, the am-
perometric response of the pine-needle-like CuO modified electrode
upon successive additions of 1.0 mM H2 O2 and 1.0 mM interfering
species (AA, UA, DA, and NaCl) were measured, and is shown in
Figure 8. Clearly it exhibits negligible responses to aforementioned
interfering species, and therefore demonstrates good selectivity to-
ward the detection of H2 O2 . After stored in air at ambient conditions
for three months, the sensitivity of the pine-needle-like CuO modified
electrode still retained 90.5% of its initial value, demonstrating good
long-term stability.

Conclusions
In summary, various Cu(OH)2 /CuO nanostructures on copper foil
substrates have been in-situ grown via an anodization approach.
The morphology evolution of Cu(OH)2 /CuO nanostructures has been
studied systematically as a function of NaOH concentration, current
density, reaction temperature and time. Nitrogen heat-treatment was
employed to convert Cu(OH)2 to CuO with unnoticeable changes Figure 8. Current response of pine-needle-like CuO modified electrode at
in their morphologies. The freestanding CuO nanostructured films −0.4 V applied potential for 1 mM H2 O2 in the presence of 1 mM ascorbic
(pine-needle-like, honeycomb-like and aggregate) were directly used acid (AA), uric acid (UA), dopamine (DA), and NaCl.

Downloaded on 2019-07-24 to IP 132.174.251.90 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 166 (10) B859-B865 (2019) B865

ORCID 21. M. J. Song, S. W. Hwang, and D. M. Whang, Talanta, 80, 1648 (2010).
22. B. J. Wang, L. Q. Luo, Y. P. Ding, D. S. Zhao, and Q. L. Zhang, Colloids Surf. B, 97,
Peng Gao https://orcid.org/0000-0002-5104-2978 51 (2012).
23. S. H. Zhou, X. Feng, H. Y. Shi, J. Chen, F. Zhang, and W. B. Song, Sens. Actuators
B: Chem., 177, 445 (2013).
24. J. P. Dong, L. X. Ren, Y. Zhang, X. L. Cui, P. F. Hu, and J. Q. Xu, Talanta, 132, 719
References (2015).
25. B. G. Mahmoud, M. Khairy, F. A. Rashwan, C. W. Foster, and C. E. Banks, RSC Adv.,
1. J. F. Ping, X. L. Mao, K. Fan, D. Y. Li, S. P. Ru, J. Wu, and Y. B. Ying, Ionics, 16, 6, 14474 (2016).
523 (2010). 26. X. F. Wu, H. Bai, J. X. Zhang, F. E. Chen, and G. Q. Shi, J. Phys. Chem. B, 109,
2. P. Gao and D. W. Liu, Sens. Actuators B: Chem., 208, 346 (2015). 22836 (2005).
3. P. Gao and D. W. Liu, RSC Adv., 5, 24625 (2015). 27. H. Ming, K. M. Pan, Y. Liu, H. T. Li, X. D. He, J. Ming, Z. Ma, and Z. H. Kang,
4. P. Gao, Y. X. Gong, N. P. Mellott, and D. L. Liu, Electrochim. Acta, 173, 31 J. Cryst. Growth, 327, 251 (2011).
(2015). 28. Y. H. Li, S. Chang, X. L. Liu, J. C. Huang, J. L. Yin, G. L. Wang, and D. X. Cao,
5. P. V. Rodionov, E. A. Alieva, E. A. Sergeeva, I. A. Veselova, and Electrochim. Acta, 85, 393 (2012).
T. N. Shekhovtsova T N, J. Anal. Chem., 71, 932 (2016). 29. L. Zaraska, N. Czopik, M. Bobruk, G. D. Sulka, J. Mech, and M. Jaskula, Electrochim.
6. T. F. Jiao, B. D. Leca-Bouvier, P. Boullanger, L. J. Blum, and A. P. A. Girard-Egrot, Acta, 104, 549 (2013).
Colloids Surf. A, 354, 284 (2010). 30. O. Sisman, N. Kilinc, and Z. Z. Ozturk, Sens. Actuators B: Chem., 236, 1118 (2016).
7. A. M. Noor, M. M. Shahid, P. Rameshkumar, and N. M. Huang, Microchim. Acta, 31. Y. Q. Wang, T. T. Jiang, A. W. Meng, H. Y. Jin, and M. H. Yu, Appl. Surf. Sci., 349,
183, 911 (2016). 636 (2015).
8. Z. H. Li, D. H. Li, K. Oshita, and S. Motomizu, Talanta, 82, 1225 (2010). 32. J. Z. Chen, J. L. Xu, S. Zhou, N. Zhao, and C. P. Wong, J. Mater. Chem. A, 3, 17385
9. P. Gao and D. W. Liu, Microchim. Acta, 182, 1231 (2015). (2015).
10. N. N. Wei, X. Xin, J. Y. Du, and J. L. Li, Biosens. Bioelectron., 26, 3602 (2011). 33. J. Wang and W. D. Zhang, Electrochim. Acta, 56, 7510 (2011).
11. Y. D. Liu, X. H. Liu, Z. P. Guo, Z. G. Hu, Z. H. Xue, and X. Q. Lu, Biosens. Bioelec- 34. X. M. Miao, R. Yuan, Y. Q. Chai, Y. T. Shi, and Y. Y. Yuan, J. Electroanal. Chem.,
tron., 87, 101 (2017). 612, 157 (2008).
12. J. Wang, Chem. Rev., 108, 814 (2008). 35. J. Bai, X. J. Bo, C. Luhana, and L. P. Guo, Electrochim. Acta, 56, 7377 (2011).
13. T. F. Tseng, Y. L. Yang, Y. J. Lin, and S. L. Lou, Sensors, 10, 6347 (2010). 36. Y. C. Li, F. Y. Huang, J. Chen, T. Mo, S. X. Li, F. Wang, S. Q. Feng, and Y. J. Li, Int.
14. U. B. Tribedi, D. Lakshminarayana, I. L. Kothari, P. B. Patel, and C. J. Panchal, Sens. J. Electrochem. Sci., 8, 6332 (2013).
Actuators B: Chem., 136, 45 (2009). 37. Y. M. Zhong, Y. C. Li, S. X. Li, S. Q. Feng, and Y. Y. Zhang, RSC Adv., 4, 40638
15. Q. Xia, Y. H. Huang, X. Lin, S. Zhu, and Y. Z. Fu, Biochem. Eng. J., 113, 1 (2014).
(2016). 38. F. G. Xu, M. Deng, G. L. Li, S. H. Chen, and L. Wang, Electrochim. Acta, 88, 59
16. K. E. Toghill and R. G. Compton, Int. J. Electrochem. Sci., 5, 1246 (2010). (2013).
17. J. Lv, C. C. Kong, Y. Xu, Z. M. Yang, X. J. Zhang, S. C. Yang, G. Meng, J. L. Bi, 39. S. H. Weng, Y. J. Zheng, C. F. Zhao, J. Z. Zhou, L. Q. Lin, Z. F. Zheng, and X. H. Lin,
J. H. Li, and S. Yang, Sens. Actuators B: Chem., 248, 630 (2017). Microchim. Acta, 180, 371 (2013).
18. S. H. Kim, A. Umar, and S. W. Hwang, Ceram. Int., 41, 9468 (2015). 40. J. Song, L. Xu, C. Y. Zhou, R. Q. Xing, Q. L. Dai, and H. W. Song, ACS Appl. Mater.
19. D. Chirizzi, M. R. Guascito, E. Filippo, C. Malitesta, and A. Tepore, Talanta, 147, Interfaces, 5, 12928 (2013).
124 (2016). 41. X. Wen, M. Long, and A. D. Tang, J. Electroanal. Chem., 785, 33 (2017).
20. J. Y. Li, S. L. Xiong, J. Pan, and Y. T. Qian, J. Phys. Chem. C, 114, 9645 (2010). 42. Y. K. Hsu, Y. C. Chen, and Y. G. Lin, Appl. Surf. Sci., 354, 85 (2015).

Downloaded on 2019-07-24 to IP 132.174.251.90 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

You might also like