You are on page 1of 10

Sensors & Actuators: B.

Chemical 345 (2021) 130384

Contents lists available at ScienceDirect

Sensors and Actuators: B. Chemical


journal homepage: www.elsevier.com/locate/snb

A synergistic promotion strategy for selective trapping and sensing of lead


(II) by oxygen-vacancy and surface modulation of MnO2 nanoflowers
Jianjun Liao a, Yang Song b, Ruyi Wang b, Yamei Zeng b, c, Hewei Si b, c, Chengjun Ge a,
Shiwei Lin b, c, *
a
Key Laboratory of Agro-Forestry Environmental Processes and Ecological Regulation of Hainan Province, School of Ecological and Environmental Sciences, Hainan
University, Haikou, 570228, China
b
School of Materials Science and Engineering, Hainan University, Haikou, 570228, China
c
State Key Laboratory of Marine Resource Utilization in South China Sea, Hainan University, Haikou, 570228, China

A R T I C L E I N F O A B S T R A C T

Keywords: Designed and fabricated electrochemical sensing materials with high electrical conductivity, sufficient number of
MnO2 active sites, and good selectivity to target analytes are highly desirable for sensitive determination of heavy metal
Phosphorization ions in water. In this work, MnO2 nanoflowers with enriched oxygen vacancies and surface phosphate ions (P-
Oxygen vacancy
MnO2-x) were prepared by a simple phosphorization process. For this novel electrode modifier, oxygen vacancies
Mn(III)/Mn(IV) cycle
Pb(II) sensing
increased the electrical conductivity, thereby improving the signal-to-noise ratio of sensors. Phosphate ions acted
as ligand molecules for trapping Pb(II) and thus enhanced the selectivity toward Pb(II), and through (PO4)3−
bridge, a fast charge-transfer channel was formed and promoting the redox cycles between Mn(III)/Mn(IV) and
Pb(II)/Pb(0). Thus, these multiple synergistic effects endowed P-MnO2-x with a high detection sensitivity of
50.11 μA μM− 1 and a low limit of detection of 0.0012 μM. More impressively, after phosphorization treatment, P-
MnO2-x showed ultra-high selectivity toward Pb(II) while did not change the stripping signals obviously toward
other common metal ions. This synergistic oxygen-vacancy and surface modulation strategy demonstrates a new
way to construct high-performance electrode materials for electroanalysis.

1. Introduction target analytes [15]. Among them, metal oxides (e.g., Fe2O3, Fe3O4,
Co3O4, MnO2, etc.) are excellent candidates due to their unique struc­
Lead(II) is a representative hazardous metal ion in the environment. tural property, large surface area, and tremendous adsorption capacity
Long-term exposure to Pb(II) ions, even at trace levels, can cause fearful [16–20]. However, metal oxides often suffer from poor conductivity,
damage to the immune and central nervous system of the human body which may affect the sensing performance of heavy metals. To solve this
[1–4]. Accordingly, it is highly desired to explore a sensitive and accu­ obstacle, conductive materials, especially graphene [21–24], carbon
rate method for the determination of Pb(II) in different media. Induc­ nanotube [25–27], and conductive polymer [28–31], are used as a
tively coupled plasma mass spectrometry (ICP-MS) [5] and atomic matrix for the loading of metal oxides. For example, Gao el al [32]. re­
absorption spectroscopy (AAS) [6] are the most accepted analytical ported AlOOH-reduced graphene oxide nanocomposites for the simul­
methods. However, their expensive and bulky instrumentation, opera­ taneous analysis of Pb(II) and Cd(II). In this hybrid, AlOOH nanoplates
tional costs, and highly trained operators are serious drawbacks. In provided a high adsorption capacity, and reduced graphene oxide
contrast to these conventional methods, electrochemical techniques offered a good conducting pathway for electrochemical detection of
have been considered promising means due to their low cost, high heavy metal ions. However, the selectivity is not ideal enough due to the
sensitivity, and easy operations [7–12]. low binding affinity toward heavy metal ions. Meanwhile, the propor­
As reported in the literature, nanomaterial modified working elec­ tion of components in the nanocomposites should be carefully controlled
trode is a common method for achieving sensitive and selective detec­ in order to prevent the aggregation problem, which may significantly
tion [13,14]. The ideal sensing materials are expected to possess high undermine its electrochemical performance. Therefore, it is highly
electrical conductivity, plentiful active sites, and good selectivity to challenging to incorporate high electrical conductivity, sufficient active

* Corresponding author at: School of Materials Science and Engineering, Hainan University, Haikou, 570228, China.
E-mail address: linsw@hainu.edu.cn (S. Lin).

https://doi.org/10.1016/j.snb.2021.130384
Received 12 May 2021; Received in revised form 17 June 2021; Accepted 27 June 2021
Available online 29 June 2021
0925-4005/© 2021 Elsevier B.V. All rights reserved.
J. Liao et al. Sensors and Actuators: B. Chemical 345 (2021) 130384

Fig. 1. (a,b) FESEM, (c) TEM, and (d) HRTEM images of P-MnO2-x. (e) STEM-EDS mapping images for Mn, O, and P elements in the selected area.

sites, and good selectivity simultaneously on single-component metal pristine MnO2 nanoflowers using a simple phosphorization process,
oxides. which significantly boosts their electrochemical sensing performance.
Surface functional groups play an important role in the selective For the design of electrode modifiers, oxygen vacancies not only
detection of heavy metals. According to the adsorb-release model [33], enhance the electrical conductivity of metal oxides but also serve as
when an adsorbent is modified with specific reactive groups, such as active sites for electrochemical reactions. Surface phosphate ions
carboxyl, amine, and thiol, some heavy metal ions would be preferen­ ((PO4)3− ) act as ligand molecules for strong binding affinity with Pb(II),
tially adsorbed by the complexation reaction [34] and thus enhance the resulting in high selectivity. At the same time, under the strong inter­
selectivity and electrochemical signal. For instance, by modification of facial electronic interaction, (PO4)3− ions accelerate the electron
multiwalled carbon nanotubes with β-cyclodextrin (βCD), Alam et al. transfer between Mn(III)/Mn(IV) cycle and Pb(II)/Pb(0) redox, thereby
[35] achieved good sensitive detection of Pb(II) in the presence of other improving the detection sensitivity. As a consequence, the resulting P-
interfering heavy metal ions, such as Zn(II), Cd(II), Cu(II), Hg(II), and Ni MnO2-x exhibited excellent sensitivity and selectivity toward Pb(II).
(II). The enhanced sensitivity can be ascribed to the selective complex­ Meanwhile, the possible sensing mechanism was systematically inves­
ation between the multiple surface functional groups on βCD and Pb(II). tigated by combining electrochemical experiments, adsorption mea­
Baghayeri et al. [36] utilized glutathione coated Fe3O4 nanoparticles for surements, X-ray photoelectron spectroscopy (XPS), and density
the simultaneous determination of Pb(II) and Cd(II) ions. They found functional theory (DFT) calculations. Finally, the anti-interference,
that the negative-charged functional groups (-COOH and -SH) at gluta­ stability, and reproducibility were evaluated, and the real water sam­
thione provided a strong affinity toward Pb(II) and Cd(II), which might ples were also detected. The proposed sensors showed good results.
be responsible for the reliability and accuracy of the sensors. Therefore,
it is a promising strategy to improve the selectivity by incorporating
specific functional groups on the electrode modifiers.
In this work, MnO2 nanoflowers with enriched oxygen vacancies and
surface phosphate ions (P-MnO2-x) were prepared by annealing the

2
J. Liao et al. Sensors and Actuators: B. Chemical 345 (2021) 130384

Fig. 2. (a) XPS survey spectrum of MnO2 and P-MnO2-x. High-resolution XPS spectra of (b) Mn 2p, (c) O 1s, and (d) P 2p resolved MnO2 and P-MnO2-x.

2. Experimental section 2.3. Eletrochemical detection

2.1. Preparation of P-MnO2-x nanoflowers An electrochemical analyzer (CHI660E, ChenHua Instruments Co.)
was used for the electrochemical measurements. P-MnO2-x nanoflowers
P-MnO2-x nanoflowers were firstly synthesized through a hydro­ modified glassy carbon electrode (P-MnO2-x/GCE) was used as working
thermal method. 0.0128 g of KMnO4 was added to 40 mL of deionized electrode to detect Pb(II) using square wave anodic stripping voltam­
water under vigorously stirring. Subsequently, the above solution was metry (SWASV). Firstly, a deposition potential of -1.0 V was applied at P-
transferred to a 100 mL Teflon-lined stainless steel autoclave and heated MnO2-x/GCE for 150 s to reduce Pb(II) to Pb(0). Then, the electro­
to 180 ◦ C for 24 h. The resulting samples were thoroughly washed with deposited Pb atoms were stripped by sweeping the potential from − 1.1 V
deionized water and dried under vacuum at 60 ◦ C. To obtain P-MnO2-x to 0 V. The optimized SWASV parameters were 15 Hz, 25 mV, and 4 mV
nanoflowers, the as-prepared samples were annealed in the N2 atmo­ for frequency, amplitude, and step potential, respectively.
sphere at 200 ◦ C for 2 h in the presence of NaH2PO2⋅H2O (2.0 g). For
comparison, the pure MnO2 was annealed in the N2 atmosphere at 200 2.4. Adsorption experiment

C for 2 h in the absence of NaH2PO2⋅H2O.
10 mg of samples were added into 10 mL NaAc-HAc buffer solution
2.2. Characterization (0.1 M, pH 5) containing Pb(II), and then shook the mixed solution at
298 k for 24 h. After reaching the adsorption equilibrium, the residual
Field emission-scanning electron microscope (FESEM, Hitachi Pb(II) concentration was determined using ICP-AES. Meanwhile, the
SU8020) and transmission electron microscopy (TEM, JEOL JEM-2010) adsorbents were separated by centrifugation at 8000 rpm for 5 min, and
were used to characterize the morphology of the samples. X-ray dried under vacuum at 60 ◦ C for XPS analysis.
photoelectron spectroscopy (XPS, AXIS SUPRA) was used to study the
chemical valence. Powder X-ray diffraction (XRD) measurements were 3. Results and discussion
performed on Bruker D8 Advance diffractometer. Theoretical calcula­
tions were performed by Vienna ab-initio simulation package (VASP) for 3.1. Characterization of the as-prepared samples
surface adsorption energy calculation. More details were presented in
Supporting Information. Fig. 1a,b present the SEM images of P-MnO2-x. It is seen that P-MnO2-
x possesses a 3D flower-like structure. The detailed structural features
could be observed from TEM. As shown in Fig. 1c, P-MnO2-x is assembled
by homogeneous 2D nanosheets. HRTEM image (Fig. 1d) indicates that

3
J. Liao et al. Sensors and Actuators: B. Chemical 345 (2021) 130384

the crystal lattice fringe is 0.253 nm, which is assigned to the (101)
crystal plane of MnO2 [37]. Moreover, homogeneous distributions of
Mn, O, and P elements are confirmed by EDS mapping images of
P-MnO2-x (Fig. 1e), demonstrating the presence of P species. XRD pat­
terns (Fig. S2) also show that all the characteristic peaks are well
indexed to δ-MnO2 (JCPDS NO.86-0666) [38], and the phosphorization
treatment does not change the crystal structure of MnO2.
XPS was conducted to explore the effect of the phosphorization
process on the surface electronic states of the samples. Fig. 2a compared
the XPS full survey of the bare MnO2 and P-MnO2-x. The elements of Mn,
O, and P could be identified in the P-MnO2-x samples, which is in accord
with the above STEM-EDS mapping results. As shown in Fig. 2b, the
binding energies of Mn 2p peaks located at 653.8 eV and 642.2 eV can be
assigned to Mn 2p1/2 and Mn 2p3/2 peaks [39], respectively, proving
that Mn(IV) is the predominant oxidation state in the pure MnO2.
Moreover, the Mn 2p3/2 peak could be deconvoluted into two typical
peaks at 642.0 eV and 641.0 eV corresponding to the existence of Mn(IV)
and Mn(III) [40], respectively. It can be seen that the content of Mn(III)
in MnO2 is relatively low (~8.2 %). However, after phosphorization
treatment, more Mn(IV) species are reduced to Mn(III), and the content
of Mn(III) in P-MnO2-x increases to 27.7 %. By means of the valence
change of Mn, the increased number of Mn(III) and Mn(IV) pairs in
P-MnO2-x may promote the electron transfer between the electrode
materials and heavy metal ions, and thus improves the detection
sensitivity. Previous studies have shown that the redox reactions based
on Co(II)/Co(III) and Fe(II)/Fe(III) cycles are beneficial for the electro­
chemical detection of heavy metal ions [41,42].
For O 1s spectra (Fig. 2c), the peaks at 532.2 eV, 530.8 eV, and 529.3
eV are ascribed to surface-absorbed water (O–H bonds), oxygen defect
sites, and lattice oxygen (Mn-O bonds), respectively [43]. In comparison
with the pristine MnO2, the intensity of oxygen vacancy sites in
P-MnO2-x becomes stronger owing to the increased concentration of
oxygen vacancies, and the concentration of oxygen vacancies for
P-MnO2-x is estimated to be about 19.4 %, which is about two times that
of the pure MnO2 (9.38 %). The oxygen vacancies are expected to
enhance the electrical conductivity and the signal-to-noise ratio of
sensors [44,45]. Besides, compared with the bare MnO2, the P 2p spectra
of P-MnO2-x (Fig. 2d) presents a characteristic peak at 133.8 eV [46],
revealing that the phosphorus exists in the form of (PO4)3− on the sur­
face of P-MnO2-x. The (PO4)3− ions may contribute to enhance the
electrochemical reactivity [39,46,47] and provide good affinity toward
Pb(II), resulting in higher sensitivity and selectivity. Therefore, the
synergistic effect of Mn(III)/Mn(IV) cycle, oxygen vacancies, and surface
(PO4)3− ions is expected to improve the redox activity, enhance the
electrical conductivity, and increase the number of adsorption sites,
leading to superior electroanalytical performance of P-MnO2-x.

3.2. Electrochemical determination of Pb(II)

The experimental conditions were firstly optimized. As shown in


Fig. S3, 0.1 M NaAc-HAc buffer solution (pH 5) was chosen as the sup­
porting electrolyte, and the optimum deposition potential and time were
selected as − 1.1 V and 150 s. Under the optimized parameters, the
SWASV responses of MnO2 and P-MnO2-x toward Pb(II) are shown in
Fig. 3a,b. For MnO2, the peak current increases linearly versus the
concentration of Pb(II) in the range of 0.2–1.4 μM. The corresponding
linearization equations is Y (μA) = − 0.49 + 23.90 X (μM), R2 = 0.998;
and the limit of detections (LOD) is calculated to be 0.025 μM (LOD =
3σ/S). While for P-MnO2-x, the peak current shows a wider linear range
Fig. 3. (a,b) SWASV responses of MnO2 and P-MnO2-x modified GCE toward Pb
of 0.02–1.4 μM. The corresponding linearization equations is Y (μA) =
(II). Insets show the corresponding calibration plots. (c) Comparison of the
0.66 + 49.56 X (μM), R2 = 0.999; and the LOD is down to 0.0012 μM
stripping peak currents of MnO2 and P-MnO2-x modified GCE toward different
metal ions with the same concentration (1 μM).
(LOD = 3σ/S). As can be seen that the sensitivity of P-MnO2-x is almost
two times that of MnO2, and the LOD of P-MnO2-x is one order of
magnitude lower than that of MnO2. Furthermore, a comparison of Pb
(II) determination with various modified electrodes in previous reports
is summarized in Table S1. It indicates that the P-MnO2-x modified

4
J. Liao et al. Sensors and Actuators: B. Chemical 345 (2021) 130384

Fig. 4. (a) CV and (b) EIS characterizations of GCE, MnO2/GCE, and P-MnO2-x/GCE.

electrode exhibited comparable performance for the detection of Pb(II).


Especially, compared with conventional Bi-based electrodes [48,49], the
P-MnO2-x modified electrodes showed good selectivity because surface
(PO4)3− ions can increase the number of active sites for Pb(II) adsorp­
tion, which will be discussed later. Besides, compared with in-situ elec­
troplated bismuth film electrodes, the P-MnO2-x modified electrodes are
cost-effective and repeatable because it avoids the use of large volumes
of Bi salt solution [50,51].
In order to investigate the effect of phosphorization treatment on the
other heavy metal ions, we compared the SWASV responses of MnO2 and
P-MnO2-x toward the same concentration of Pb(II), Cd(II), Ni(II), Co(II),
Cu(II), As(III), Hg(II), and Zn(II). As shown in Fig. 3c, it is observed that
the pure MnO2 is suitable for the simultaneous detection of Pb(II) and Cd
(II) in water since the response current values of these two ions are very
close. However, after phosphorization treatment, the peak current of P-
MnO2-x exhibits a slight increase toward Cd(II), and the current change
rate is about 23.7 %. Surprisingly, P-MnO2-x shows a distinct current
change toward Pb(II). The peak current of P-MnO2-x toward Pb(II) is
48.94 μA, which is almost two times that of pure MnO2 (24.2 μA).
Meanwhile, the peak current of Pb(II) on P-MnO2-x is also much larger
than other metal ions, even about 25 times difference could be found
when used for Hg(II) detection. The ultra-highly selectivity toward Pb
(II) could be ascribed to the synergistic effect of Mn(III)/Mn(IV) cycle,
oxygen vacancy sites, and surface (PO4)3− ions on P-MnO2-x, which will
be discussed below.

3.3. Evidence for improved electrochemical performance

Fig. 4a represents the different cyclic voltammograms (CVs) of the


bare GCE, MnO2/GCE, and P-MnO2-x/GCE in the presence of 5 mM [Fe
(CN)6]3− /4− redox probe in 0.1 M KCl solution. Compared with the bare
GCE, the peak current and peak separation show a slight decrease after
modification with MnO2 or P-MnO2-x onto the GCE surface, indicating
the successful construction of the electrochemical sensing interface.
Moreover, compared to MnO2, the P-MnO2-x modified electrode shows a
slight increase of redox peaks current, mainly due to enhanced electrical
conductivity and a large specific surface area. This speculation was also
verified by the active electrode area and electrochemical impedance
spectroscopy (EIS) measurements. As showed in Fig. S3, the active
Fig. 5. (a) The plot of adsorption capacities of Pb(II) by MnO2 and P-MnO2-x
electrode areas of MnO2 and P-MnO2-x modified GCE were evaluated to
versus adsorption time. (b) XPS spectra of Pb 4f after Pb(II) adsorption on MnO2
be 0.068 cm2 and 0.1063 cm2, respectively. The increased active elec­ and P-MnO2-x, and inset compares the atomic ratio of Pb(II) adsorption on
trode area of P-MnO2-x is expected to offer adequate active sites for MnO2 and P-MnO2-x.
electrochemical reaction. Fig. 4b displays the EIS curves of various
modified electrodes in the frequency range from 0.1 Hz to 100 kHz. In a
GCE is about 32.8 Ω. However, after modification with MnO2, the Ret
typical EIS Nyquist plot, the semicircle diameter in a higher frequency
value increases to about 2042 Ω, which is ascribed to that the poor
region equals the electron-transfer resistance (Ret), reflecting the the
conductivity of MnO2. Furthermore, after P-MnO2-x is modified on the
electrode’s interface properties. It can be found that the Rct value of bare

5
J. Liao et al. Sensors and Actuators: B. Chemical 345 (2021) 130384

Table 1 atom prefers to occupy the vacancy surrounded by the O atoms and
Adsorption energies of Pb on MnO2 and P-MnO2. binds with the O atoms around it [19,53]. Hence, as shown in Fig. S4a,
Adsorption Energy (eV) there are eight most favorable adsorption sites on the surface of MnO2
(marked as S on Fig. S4a), but when the (PO4)3− ions are introduced,
Pb/MnO2 − 4.46
Pb/P-MnO2 (Sa) − 6.06 three new adsorption sites could be produced (marked as Sa, Sb, and Sc
Pb/P-MnO2 (Sb) − 6.25 on Fig. S4b, respectively), whereas there are eleven adsorption sites on
Pb/P-MnO2 (Sc) − 6.50 the surface of P-MnO2. The corresponding possible adsorption configu­
rations are shown in Fig. 5. In the Pb/MnO2 adsorption system (Fig. 5a),
Pb is coordinated with three O atoms with the Pb-O bond lengths of
GCE, the Ret value distinctly decreases to 934.2 Ω, which indicates that
2.267, 2.299, and 2.340 Å, respectively. While for the Pb/P-MnO2
the faster electron transfer and higher conductivity are induced by ox­
adsorption system, Pb not only binds with the surface oxygen atoms on
ygen vacancy.
Fig. 5a shows the plot of adsorption amount of Pb(II) versus
adsorption time. As the adsorption time increases, the adsorption ca­
pacity increases until the adsorption reaches the equilibrium phase, and
the equilibrium sorption is almost achieved within 10 min. The exper­
imental results show that P-MnO2-x has a better adsorption effect on Pb
(II). Furthermore, the adsorption of Pb(II) on MnO2 and P-MnO2-x were
probed by XPS (Fig. 5b). Compared with the binding energies of Pb 4f on
MnO2, the peaks of Pb 4f absorbed on P-MnO2-x show a clear shift to the
low binding energies of 0.8 eV, suggesting that the binding between Pb
(II) and P-MnO2-x is stronger than Pb(II) with MnO2 [15,52]. The dif­
ference in binding energy can be ascribed to the strong interaction be­
tween Pb(II) and the surface (PO4)3− ions on P-MnO2-x, as proved by
DFT calculations. Beyond that, the amount of Pb(II) adsorbed on
P-MnO2-x (0.55 %, atomic ratio) is three times that of MnO2 (0.18 %,
atomic ratio). These results also demonstrate that the adsorption ability
has a positive correlation with the electrochemical sensing performance.
To better understand the positive effect of surface phosphorization
treatment on the improved electrochemical performance of P-MnO2-x,
DFT calculations were performed to explore the adsorption behaviors of
Pb(II) on P-MnO2-x at the atomic level. Here, it should be noted that
because we only investigate the role of (PO4)3− ions, the oxygen vacancy Fig. 7. Charge density difference of (PO4)3− adsorption on MnO2. The yellow
and cyan colors represent the increase and decrease of electron density,
is not introduced into our slab model in order to increase the compu­
respectively. The isovalues for plotting is 0.003 e/Å3 (For interpretation of the
tational speed and efficiency. Fig. S1 shows the optimized geometry
references to colour in this figure legend, the reader is referred to the web
model of MnO2 and P-MnO2-x. Our previous study indicates that metal version of this article).

Fig. 6. Optimized adsorption configurations of Pb on (a) MnO2 and (b,c,d) P-MnO2.

6
J. Liao et al. Sensors and Actuators: B. Chemical 345 (2021) 130384

Scheme 1. Schematic diagram illustrating the trapping and sensing strategy of P-MnO2-x toward Pb(II) in water.

MnO2 but also could absorb on the hollow site created by (PO4)3− ions.
The optimal adsorption states of Pb/P-MnO2 are depicted in Fig. 5b–d. It
indicated that P-MnO2 could provide more adsorption sites for heavy
metal ions and thus enhance its electrochemical stripping signals.
In addition to the adsorption site, the adsorption energy also in­
creases after surface phosphorization treatment. Table 1 lists the
calculated adsorption energies corresponding to the optimized adsorp­
tion structures shown in Fig. 6. Clearly, Pb has the lowest adsorption
energy on MnO2 (− 4.46 eV), whereas the adsorption energies increased
to − 6.06/− 6.25/− 6.50 eV after the introduction of (PO4)3− ions. These
calculation results support that both the adsorption site and adsorption
energy have increased because of the existing surface (PO4)3− ions on
MnO2. Besides, Fig. 7 shows the charge density difference of (PO4)3−
adsorption on MnO2. The yellow and the cyan part represent the accu­
mulation and depletion electrons, respectively. It can be seen that the
charge reconstruction mainly occurs at the (PO4)3− /MnO2 interface
region, and the electrons are mostly localized around (PO4)3− ions,
which indicates that surface (PO4)3− ions form an interfacial built-in
field on the (PO4)3− /MnO2 interface region. The formed interfacial
built-in field may lead to a fast electron transfer between Mn(III)/Mn
(IV) cycle and Pb(II)/Pb(0) redox [54,55], promoting the electro­
Fig. 8. Interference study of other ions in Pb(II) detection. I0 and I represent the
chemical sensing performance.
stripping currents of Pb(II) in the absence and presence of interfering species,
respectively.
3.4. Possible sensing mechanism
between Mn(III)/Mn(IV) cycle and Pb(II)/Pb(0) redox. The reaction can
According to the experimental results and analysis above, adsorption be described as follows:
capacity and electroanalytical reactivity of nano-metal oxides are
Mn(III)sur + Pb(II)abs → Mn(IV)sur + Pb(0)abs (Pre-concentration)
beneficial for achieving high detection sensitivity and selectivity.
Scheme 1 illustrates the detection mechanism of Pb(II) ions on the Mn(IV)sur + Pb(0)abs → Mn(III)sur + Pb(II)abs (Stripping)
surface of P-MnO2-x based on the trapping and sensing strategy. Firstly,
the surface (PO4)3− ions can selectively trap the target Pb(II) ions and
bring them close to the surface of P-MnO2-x even when other metal ions
coexisted in water. Under the mediating effect of Mn(III)/Mn(IV) cycle, 3.5. Interference measurements
the adsorbed Pb(II) can be reduced to Pb(0) during the pre-
concentration process, thereby promoting the content of Pb(0)abs on The interference-free detection of heavy metal ions is an important
the electrode surface. And then, Pb(0)abs is oxidized to Pb(II) During the character for the developed sensor. Here, the tolerance ability is defined
stripping process, resulting in enhanced electrochemical stripping sig­ as the peak current changes keep the relative error of 10 % in the ex­
nals at SWASV. Meanwhile, the Mn(IV)sur is reduced to Mn(III)sur, and istence of interfering species. As shown in Fig. 8, 5-fold Fe(III), 1-fold Al
the surface Mn(III)/Mn(IV) redox cycle is completed. Obviously, the Mn (III), Ba(II), and Mg(II), 2-fold Ca(II), Cu(II), and Cd(II), 10-fold Cl− ,
(III)/Mn(IV) cycle serves as the electron mediator for effective electron SO2−4 , and NO3 could satisfy the tolerance ratio. The results suggest that

transfer in electrochemical detection. The surface (PO4)3− ions act as the developed sensor can be used to analyze Pb(II) in the real sample.
active capture sites for the selective adsorption of Pb(II) in water and
serve as a charge-transfer channel to promote the electron transfer

7
J. Liao et al. Sensors and Actuators: B. Chemical 345 (2021) 130384

3.7. Analysis of real water samples

To evaluate the practicability of P-MnO2-x/GCE, the determination of


Pb(II) in seawater and tap water was carried out using the standard
addition method. Before measurements, the water samples were diluted
with 0.1 M NaAc buffer solution (pH 5.0) in the ratio of 1:2 (v:v). The
determined Pb(II) concentrations and the recoveries are summarized in
Table 2. The recoveries obtained are varied from 95.75%–98.75%,
revealing that P-MnO2-x/GCE has practical application potential.

4. Conclusion

In summary, we developed a synergistic promotion strategy for se­


lective trapping and sensing of Pb(II) by oxygen-vacancy and phosphate-
ion modulation of MnO2 nanoflowers. By the introduction of oxygen
vacancies and phosphate ions, P-MnO2-x displayed good electronic
conductivity, sufficient number of active sites, and impressive selective
adsorption capacity toward Pb(II). Especially, through (PO4)3− bridge,
the interfacial electronic effects accelerate the electron transfer of Mn
(III)/Mn(IV) cycle and promote the redox reaction of Pb(II)/Pb(0),
further improving the electrochemical response. In addition, the P-
MnO2-x modified electrodes showed good stability and reproducibility.
Analysis of real water samples demonstrated the great potential for
electrochemical detection of Pb(II) in practical applications. This
oxygen-vacancy and surface modulation strategy can also be extended to
the selective detection of other heavy metal ions by the rational func­
tionalization with proper ligand molecules. Furthermore, many research
efforts have been made to develop the miniaturized analytical system as
the increasing requirements for point-of-care testing (POCT). After
modification with the suitable sensing materials, different low-cost,
portable, and sensitive sensors systems could be easily realized.

CRediT authorship contribution statement

Jianjun Liao: Conceptualization, Writing-original draft, Project


administration. Yang Song: Investigation, Methodology, Data curation.
Ruyi Wang: Investigation, Methodology, Data curation. Yamei Zeng:
Theoretical calculation & analysis. Hewei Si: Visualization, Investiga­
tion. Chengjun Ge: Writing-review & editing. Shiwei Lin: Project
Fig. 9. (a) Reproducibility on the same electrode ten times repeatedly. (b) administration, Funding acquisition.
Electrode stability in response to 0.4, 1.0, and 1.6 μM Pb(II) over ten days.

Declaration of Competing Interest


Table 2
Analytical performance for Pb(II) determination in different real samples (Each The authors report no declarations of interest.
sample was tested for three times).
Sample Added (μM) Found (μM) Recovery (%) RSD (%) Acknowledgments
Seawater 0.4 0.395 ± 0.006 98.75 1.52
0.6 0.578 ± 0.009 96.33 1.52 This work is supported by the National Natural Science Foundation
Tap water 0.4 0.383 ± 0.015 95.75 3.91 of China (21866012, 61764003), Basic and Applied Basic Research
0.6 0.581 ± 0.012 96.83 2.06 Program of Hainan Province (2019RC023), Hainan Provincial Key
Research and Development Program (ZDYF2020222), and Scientific
Research Foundation of Hainan University (kyqd1659).
3.6. Stability and reproducibility

The stability and reproducibility were evaluated by the relative Appendix A. Supplementary data
standard deviation (RSD). As presented in Fig. 9a, the repeatability was
checked by repeatedly determining 0.8 μM Pb(II) using the same elec­ Supplementary material related to this article can be found, in the
trode under the optimized experimental conditions. The RSD of 0.96 % online version, at doi:https://doi.org/10.1016/j.snb.2021.130384.
is obtained for ten times continuous detections, indicating good
repeatability. Similarly, the stability was also checked by comparing the References
peak currents of 0.4, 1.0, and 1.6 μM Pb(II) under consecutive ten-day
[1] D. Pan, Y. Wang, Z. Chen, T. Lou, W. Qin, Nanomaterial/ionophore-based electrode
measurements (Fig. 9b). The RSD values are below 3%, indicating for anodic stripping voltammetric determination of lead: an electrochemical
excellent stability. sensing platform toward heavy metals, Anal. Chem. 81 (2009) 5088–5094.
[2] Y. Lu, X. Liang, C. Niyungeko, J. Zhou, J. Xu, G. Tian, A review of the identification
and detection of heavy metal ions in the environment by voltammetry, Talanta 178
(2018) 324–338.

8
J. Liao et al. Sensors and Actuators: B. Chemical 345 (2021) 130384

[3] A. Nsabimana, S.A. Kitte, T.H. Fereja, M.I. Halawa, W. Zhang, G. Xu, Recent [29] Q. Bao, G. Li, Z. Yang, P. Pan, J. Liu, J. Chang, et al., Electrochemical performance
developments in stripping analysis of trace metals, Curr. Opin. Electrochem. 17 of a three-layer electrode based on Bi nanoparticles, multi-walled carbon nanotube
(2019) 65–71. composites for simultaneous Hg(II) and Cu(II) detection, Chin. Chem. Lett. 331
[4] S. Mukherjee, S. Bhattacharyya, K. Ghosh, S. Pal, A. Halder, M. Naseri, et al., (2020) 2752–2756.
Sensory development for heavy metal detection: a review on translation from [30] M.H. Motaghedifard, S.M. Pourmortazavi, S. Mirsadeghi, Selective and sensitive
conventional analysis to field-portable sensor, Trends Food Sci. Technol. 109 detection of Cr(VI) pollution in waste water via polyaniline/sulfated zirconium
(2021) 674–689. dioxide/multi walled carbon nanotubes nanocomposite based electrochemical
[5] G. Álvarez-Llamas, Mad.R. Fernández de laCampa, A. Sanz-Medel, ICP-MS for sensor, Sens. Actuators B Chem. 327 (2020), 128882.
specific detection in capillary electrophoresis, TrAC Trends Anal. Chem. 24 (2005) [31] N. Wang, W. Zhao, Z. Shen, S. Sun, H. Dai, H. Ma, et al., Sensitive and selective
28–36. detection of Pb(II) and Cu(II) using a metal-organic framework/polypyrrole
[6] Y. Cai, Speciation and analysis of mercury, arsenic, and selenium by atomic nanocomposite functionalized electrode, Sens. Actuators B Chem. 304 (2020),
fluorescence spectrometry, TrAC Trends Anal. Chem. 19 (2000) 62–66. 127286.
[7] I. Rutyna, M. Korolczuk, Determination of lead and cadmium by anodic stripping [32] C. Gao, X.-Y. Yu, R.-X. Xu, J.-H. Liu, X.-J. Huang, AlOOH-reduced graphene oxide
voltammetry at bismuth film electrodes following double deposition and stripping nanocomposites: one-pot hydrothermal synthesis and their enhanced
steps, Sens. Actuators B Chem. 204 (2014) 136–141. electrochemical activity for heavy metal ions, ACS Appl. Mater. Interfaces 4 (2012)
[8] L. Xiao, H. Xu, S. Zhou, T. Song, H. Wang, S. Li, et al., Simultaneous detection of Cd 4672–4682.
(II) and Pb(II) by differential pulse anodic stripping voltammetry at a nitrogen- [33] J. Wei, S.-S. Li, Z. Guo, X. Chen, J.-H. Liu, X.-J. Huang, Adsorbent assisted in situ
doped microporous carbon/Nafion/bismuth-film electrode, Electrochim. Acta 143 electrocatalysis: an ultra-sensitive detection of As(III) in water at Fe3O4 nanosphere
(2014) 143–151. densely decorated with au nanoparticles, Anal. Chem. 88 (2016) 1154–1161.
[9] T. Yang, R. Yu, Y. Yan, H. Zeng, S. Luo, N. Liu, et al., A review of ratiometric [34] D. Zhang, L. Wang, H. Zeng, B. Rhimi, C. Wang, Novel polyethyleneimine
electrochemical sensors: from design schemes to future prospects, Sens. Actuators B functionalized chitosan–lignin composite sponge with nanowall-network structures
Chem. 274 (2018) 501–516. for fast and efficient removal of Hg(II) ions from aqueous solution, Environ. Sci.
[10] J. Liao, Z. Tao, S. Lin, Theoretical and experimental insights into the Nano 7 (2020) 793–802.
electrochemical heavy metal ion sensing with nonconductive nanomaterials, Curr. [35] A.U. Alam, M.M.R. Howlader, N.-X. Hu, M.J. Deen, Electrochemical sensing of lead
Opin. Electrochem. 17 (2019) 1–6. in drinking water using β-cyclodextrin-modified MWCNTs, Sens. Actuators B
[11] M. Yang, P.-H. Li, S.-H. Chen, X.-Y. Xiao, X.-H. Tang, C.-H. Lin, et al., Nanometal Chem. 296 (2019), 126632.
oxides with special surface physicochemical properties to promote electrochemical [36] M. Baghayeri, A. Amiri, B. Maleki, Z. Alizadeh, O. Reiser, A simple approach for
detection of heavy metal ions, Small 16 (2020), 2001035. simultaneous detection of cadmium(II) and lead(II) based on glutathione coated
[12] X. Wang, Y. Qi, Y. Shen, Y. Yuan, L. Zhang, C. Zhang, et al., A ratiometric magnetic nanoparticles as a highly selective electrochemical probe, Sens. Actuators
electrochemical sensor for simultaneous detection of multiple heavy metal ions B Chem. 273 (2018) 1442–1450.
based on ferrocene-functionalized metal-organic framework, Sens. Actuators B [37] L. Wang, S. Guan, Y. Weng, S.-M. Xu, H. Lu, X. Meng, et al., Highly efficient
Chem. 310 (2020), 127756. vacancy-driven photothermal therapy mediated by ultrathin MnO2 nanosheets,
[13] M. Li, H. Gou, I. Al-Ogaidi, N. Wu, Nanostructured sensors for detection of heavy ACS Appl. Mater. Interfaces 11 (2019) 6267–6275.
metals: a review, ACS Sustain. Chem. Eng. 1 (2013) 713–723. [38] Y. Zhao, C. Chang, F. Teng, Y. Zhao, G. Chen, R. Shi, et al., Defect-engineered
[14] C.-S. Liu, J. Li, H. Pang, Metal-organic framework-based materials as an emerging ultrathin δ-MnO2 nanosheet arrays as bifunctional electrodes for efficient overall
platform for advanced electrochemical sensing, Coord. Chem. Rev. 410 (2020), water splitting, Adv. Energy Mater. 7 (2017), 1700005.
213222. [39] Y. Zhang, S. Deng, G. Pan, H. Zhang, B. Liu, X.-L. Wang, et al., Introducing oxygen
[15] S.-S. Li, W.-Y. Zhou, M. Jiang, L.-N. Li, Y.-F. Sun, Z. Guo, et al., Insights into diverse defects into phosphate ions intercalated manganese dioxide/vertical multilayer
performance for the electroanalysis of Pb(II) on Fe2O3 nanorods and hollow graphene arrays to boost flexible zinc ion storage, Small Methods 4 (2020),
nanocubes: toward analysis of adsorption sites, Electrochim. Acta 288 (2018) 1900828.
42–51. [40] L. Ren, D. Zhou, J. Wang, T. Zhang, Y. Peng, G. Chen, Biomaterial-based flower-like
[16] X.-Z. Yao, Z. Guo, Q.-H. Yuan, Z.-G. Liu, J.-H. Liu, X.-J. Huang, Exploiting MnO2@ carbon microspheres for rapid adsorption of amoxicillin from wastewater,
differential electrochemical stripping behaviors of Fe3O4 nanocrystals toward J. Mol. Liq. 309 (2020), 113074.
heavy metal ions by crystal cutting, ACS Appl. Mater. Interfaces 6 (2014) [41] S.-S. Li, W.-Y. Zhou, M. Jiang, Z. Guo, J.-H. Liu, L. Zhang, et al., Surface Fe(II)/Fe
12203–12213. (III) cycle promoted ultra-highly sensitive electrochemical sensing of arsenic(III)
[17] Z.-G. Liu, Y.-F. Sun, W.-K. Chen, Y. Kong, Z. Jin, X. Chen, et al., Facet-dependent with dumbbell-like Au/Fe3O4 nanoparticles, Anal. Chem. 90 (2018) 4569–4577.
stripping behavior of Cu2O microcrystals toward lead ions: a rational design for the [42] X.-L. Cheng, Q.-Q. Xu, S.-S. Li, J. Li, Y. Zhou, Y. Zhang, et al., Oxygen vacancy
determination of lead ions, Small 11 (2015) 2493–2498. enhanced Co3O4/ZnO nanocomposite with small sized and loose structure for
[18] J. Liao, J. Zhang, C.-Z. Wang, S. Lin, Electrochemical and density functional theory sensitive electroanalysis of Hg(II) in subsidence area water, Sens. Actuators B
investigation on the differential behaviors of core-ring structured NiCo2O4 Chem. 326 (2020), 128967.
nanoplatelets toward heavy metal ions, Anal. Chim. Acta 1022 (2018) 37–44. [43] F. Raza, X. Ni, J. Wang, S. Liu, Z. Jiang, C. Liu, et al., Ultrathin honeycomb-like
[19] J. Liao, F. Yang, C.-Z. Wang, S. Lin, The crystal facet-dependent electrochemical MnO2 on hollow carbon nanofiber networks as binder-free electrode for flexible
performance of TiO2nanocrystals for heavy metal detection: Theoretical prediction symmetric all-solid-state supercapacitors, J. Energy Storage 30 (2020), 101467.
and experimental proof, Sens. Actuators B Chem. 271 (2018) 195–202. [44] A. Hazra, K. Dutta, B. Bhowmik, P.P. Chattopadhyay, P. Bhattacharyya, Room
[20] R. Hu, X. Zhang, K.-N. Chi, T. Yang, Y.-H. Yang, Bifunctional MOFs-based temperature alcohol sensing by oxygen vacancy controlled TiO2 nanotube array,
ratiometric electrochemical sensor for multiplex heavy metal ions, ACS Appl. Appl. Phys. Lett. 105 (2014), 081604.
Mater. Interfaces 12 (2020) 30770–30778. [45] Y. Zhong, W. Li, X. Zhao, X. Jiang, S. Lin, Z. Zhen, et al., High-response room-
[21] S. Xiong, B. Yang, D. Cai, G. Qiu, Z. Wu, Individual and simultaneous stripping temperature NO2 sensor and ultrafast humidity sensor based on SnO2 with rich
voltammetric and mutual interference analysis of Cd2+, Pb2+ and Hg2+ with oxygen vacancy, ACS Appl. Mater. Interfaces 11 (2019) 13441–13449.
reduced graphene Oxide-Fe3O4 nanocomposites, Electrochim. Acta 185 (2015) [46] Y. Zeng, Z. Lai, Y. Han, H. Zhang, S. Xie, X. Lu, Oxygen-vacancy and surface
52–61. modulation of ultrathin nickel cobaltite nanosheets as a high-energy cathode for
[22] S. Lee, J. Oh, D. Kim, Y. Piao, A sensitive electrochemical sensor using an iron advanced Zn-Ion batteries, Adv. Mater. 30 (2018), 1802396.
oxide/graphene composite for the simultaneous detection of heavy metal ions, [47] S. Liu, Y. Yin, D. Ni, K.S. Hui, K.N. Hui, S. Lee, et al., Phosphorous-containing
Talanta 160 (2016) 528–536. oxygen-deficient cobalt molybdate as an advanced electrode material for
[23] L. Ma, X. Zhang, M. Ikram, M. Ullah, H. Wu, K. Shi, Controllable synthesis of an supercapacitors, Energy Storage Mater. 19 (2019) 186–196.
intercalated ZIF-67/EG structure for the detection of ultratrace Cd2+, Cu2+, Hg2+ [48] I. Švancara, C. Prior, S.B. Hočevar, J. Wang, A decade with bismuth-based
and Pb2+ ions, Chem. Eng. J. 395 (2020), 125216. electrodes in electroanalysis, Electroanalysis 22 (2010) 1405–1420.
[24] M. Baghayeri, M. Ghanei-Motlagh, R. Tayebee, M. Fayazi, F. Narenji, Application [49] V. Jovanovski, S.B. Hočevar, B. Ogorevc, Bismuth electrodes in contemporary
of graphene/zinc-based metal-organic framework nanocomposite for electroanalysis, Curr. Opin. Electrochem. 3 (2017) 114–122.
electrochemical sensing of As(III) in water resources, Anal. Chim. Acta 1099 [50] E. Svobodova-Tesarova, L. Baldrianova, M. Stoces, I. Svancara, K. Vytras, S.
(2020) 60–67. B. Hocevar, et al., Antimony powder-modified carbon paste electrodes for
[25] A.M. Ashrafi, S. Cerovac, S. Mudrić, V. Guzsvány, L. Husáková, I. Urbanová, et al., electrochemical stripping determination of trace heavy metals, Electrochim. Acta
Antimony nanoparticle-multiwalled carbon nanotubes composite immobilized at 56 (2011) 6673–6677.
carbon paste electrode for determination of trace heavy metals, Sens. Actuators B [51] M. Gich, C. Fernandez-Sanchez, L.C. Cotet, P. Niu, A. Roig, Facile synthesis of
Chem. 191 (2014) 320–325. porous bismuth-carbon nanocomposites for the sensitive detection of heavy metals,
[26] X. Xuan, J.Y. Park, A miniaturized and flexible cadmium and lead ion detection J. Mater. Chem. A 1 (2013) 11410–11418.
sensor based on micro-patterned reduced graphene oxide/carbon nanotube/ [52] S.-S. Li, M. Jiang, T.-J. Jiang, J.-H. Liu, Z. Guo, X.-J. Huang, Competitive
bismuth composite electrodes, Sens. Actuators B Chem. 255 (2018) 1220–1227. adsorption behavior toward metal ions on nano-Fe/Mg/Ni ternary layered double
[27] P.N.D. Duoc, N.H. Binh, T.V. Hau, C.T. Thanh, P.V. Trinh, N.V. Tuyen, et al., hydroxide proved by XPS: evidence of selective and sensitive detection of Pb(II),
A novel electrochemical sensor based on double-walled carbon nanotubes and J. Hazard. Mater. 338 (2017) 1–10.
graphene hybrid thin film for arsenic(V) detection, J. Hazard. Mater. 400 (2020),
123185.
[28] K.B.R. Teodoro, F.M. Shimizu, V.P. Scagion, D.S. Correa, Ternary nanocomposites
based on cellulose nanowhiskers, silver nanoparticles and electrospun nanofibers:
use in an electronic tongue for heavy metal detection, Sens. Actuators B Chem. 290
(2019) 387–395.

9
J. Liao et al. Sensors and Actuators: B. Chemical 345 (2021) 130384

[53] J. Liao, J. Zhang, C.-Z. Wang, S. Lin, Electrochemical and density functional theory Yamei Zeng is an postgraduate student in the School of Materials Science and Engineer­
investigation on the differential behaviors of core-ring structured NiCo2O4 ing, Hainan University. Her research interest includes computational Materials Science.
nanoplatelets toward heavy metal ions, Anal. Chim. Acta 1022 (2018) 37–44.
[54] Z. Zhang, M. Wang, H. Zhou, F. Wang, Surface sulfate ion on CdS catalyst enhances
Hewei Si is an PhD student in the School of Materials Science and Engineering, Hainan
syngas generation from Biopolyols, J. Am. Chem. Soc. 143 (2021) 6533–6541.
University. Her research interest includes chemical sensors.
[55] W. Li, L. Jin, F. Gao, H. Wan, Y. Pu, X. Wei, et al., Advantageous roles of phosphate
decorated octahedral CeO2 {111}/g-C3N4 in boosting photocatalytic CO2
reduction: charge transfer bridge and Lewis basic site, Appl. Catal. B 294 (2021), Chengjun Ge is a professor in the School of Ecology and Environment, Hainan University.
120257. His current research interests include environmental tonicology and monitoring
technology.
Jianjun Liao is an associate professor in the School of Ecological and Environmental
Sciences, Hainan University. His current research interests include chemical sensors and Shiwei Lin is a professor in the School of Materials and Chemical Engineering in Hainan
intelligent systems. University. He received his Phd degree in Electrical and Electronic Engineering from
University of Manchester in the U.K. in 2006, Then he conducted two-year postdoctoral
research in microelectronics at University of Manchester. His current research interests
Yang Song is an undergraduate student in the School of Materials Science and Engi­
include photocatalytic materials and chemical sensors.
neering, Hainan University. His research interest includes nanomaterials.

Ruyi Wang is an undergraduate student in the School of Materials Science and Engi­
neering, Hainan University. Her research interest includes nanomaterials.

10

You might also like