Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Sensors & Actuators: B.

Chemical 325 (2020) 128783

Contents lists available at ScienceDirect

Sensors and Actuators: B. Chemical


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

MOF-derived ZnFe2O4/(Fe-ZnO) nanocomposites with enhanced acetone


sensing performance
Ensi Cao a, b, *, Zhaoqing Guo a, Guoqing Song a, Yongjia Zhang a, Wentao Hao a, Li Sun a,
Zhongquan Nie b
a
College of Physics and Optoelectronics, Taiyuan University of Technology, Taiyuan 030024, PR China
b
Key Lab of Advanced Transducers and Intelligent Control System, Ministry of Education, Taiyuan University of Technology, Taiyuan 030024, PR China

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

Keywords: To further improve the acetone sensing performance of MOF-derived ZnO nanoparticles, the doping of Fe into the
ZnO ZnO lattice and the heterostructure formation with ZnFe2O4 nanoparticles were simultaneously achieved by a
Zn solution-based method via the hydrolyzation of Zn-based MOF-5 in Fe(NO3)3 aqueous solution. The obtained
Fe2O4
ZnFe2O4/(Fe-ZnO) nanocomposites (the mole ratio of Fe/Zn = 0, 0.05, 0.10, 0.15, 0.20, 0.25 and 0.30) were
MOF
characterized and confirmed by XRD, TEM, SAED, Element mapping, SEM, XPS and BET. Compared with the
Nanocomposite
Acetone sensing pristine ZnO nanoparticles, better acetone sensing characteristics in terms of gas response, sensitivity, stability,
XPS reproducibility and selectivity were achieved by the ZnFe2O4/(Fe-ZnO) nanocomposites at lower prime working
temperatures, among which the nanocomposites with Fe/Zn = 0.15 exhibited the best comprehensive perfor­
mance due to its largest BET surface area, abundance of adsorption sites for oxygen and acetone molecules, as
well as the high and wide potential barrier at the interfaces.

1. Introduction structure of MOFs, and the high specific surface area makes the
MOF-derived MOSs exhibit better gas sensing performance. The ZnO
Acetone, a kind of extensively used hazardous chemicals in various nanocages prepared by the thermal decomposition of Zn-based MOF-5
industrial processes, can give rise to central nervous system anesthesia, showed good sensitivity and selectivity towards acetone owing to the
skin and eyes irritation, and nausea. Besides, the acetone in human more exposed active sites and large specific surface area, which
breath exhaust can be utilized as a critical biomarker for the noninvasive benefited the diffusion and reaction of acetone molecules on the surface
diagnosis of diabetes in early stage. Consequently, it is of significant [8]. To further improve the acetone sensing performance of
importance to make high-performance acetone sensors for the sake of MOF-derived ZnO nanomaterials, the introduction of noble metal Au
environment and human health [1,2]. was frequently utilized due to its catalytic activity and chemical sensi­
Zinc oxide (ZnO) has been extensively studied as acetone sensing or tization effect [9,10]. Additionally, the heterostructure of ZnFe2O4/ZnO
modifying material due to its high chemical stability, electron mobility, was widely studied owing to the electrical sensitization effect of
and structure tunability [3–6]. However, the high working temperature, ZnFe2O4, which can be beneficial for carrier transfer and increase in
low sensitivity, and poor selectivity still need to be improved to satisfy depletion layer width [2,11–13]. In these studies, the heterostructure
the requirement of practical application. Metal oxide frameworks was reported to be achieved by the thermal annealing of MOFs, such as
(MOFs), consisted of metal cations or metal cation aggregates linked by Prussian Blue analogue of Zn3[Fe(CN)6]2⋅xH2O [2] and Fe(III)-doped
organic ligand molecules, are promising materials for application in the MOF-5 [11], or by immersing the as-prepared ZnO in Fe-contained
sensing field due to the high porosity, specific surface area, and tena­ aqueous solution, such as FeSO4 [12] and Fe(NO3)3 [13].
bility of morphology [7]. However, the high electrical resistance hinders Herein, the synergistic effect of doping and heterostructure forma­
MOFs from direct application as electrical sensing material. While the tion on the acetone sensing performance of MOF-derived ZnO nano­
thermal annealing of MOF precursors could result in metal oxides particles was studied by comparing the acetone sensing characteristics
semiconductors (MOSs) preserving the morphology and porous of pristine ZnO nanoparticles and ZnFe2O4/(Fe-ZnO) nanocomposites.

* Corresponding author at: College of Physics and Optoelectronics, Taiyuan University of Technology, Taiyuan 030024, PR China.
E-mail address: caoensi@163.com (E. Cao).

https://doi.org/10.1016/j.snb.2020.128783
Received 29 February 2020; Received in revised form 4 June 2020; Accepted 18 August 2020
Available online 21 August 2020
0925-4005/© 2020 Elsevier B.V. All rights reserved.
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

Fig. 1. XRD patterns of (a): as-prepared MOF-5 and (b): MOF-5 after immersion in Fe(NO3)3 aqueous solution with x = 0.15.

The pristine ZnO nanoparticles were prepared by the thermal annealing ZnFe2O4/(Fe-ZnO) nanocomposites.
of Zn-based MOF-5 synthesized by a conventional solvothermal method,
while a new route to prepare ZnFe2O4/(Fe-ZnO) nanocomposites with 2.2. Characterization of sensing materials
different mole ratios of Fe/Zn was provided by immersing Zn-based
MOF-5 in Fe(NO3)3 aqueous solution before annealing treatment. The The crystalline structure of as-obtained pristine ZnO nanoparticles
ZnFe2O4/(Fe-ZnO) nanocomposites were confirmed and characterized and ZnFe2O4/(Fe-ZnO) nanocomposites was measured by X-ray
by XRD, TEM, SAED, Element mapping, SEM, XPS and BET, and the diffractometer (D/max 2500, Rigaku Corporation, Japan) using Cu Kα
origin for the best acetone sensing performance of ZnFe2O4/(Fe-ZnO) radiation. The morphology was examined by field emission scanning
nanocomposite with Fe/Zn = 0.15 was analyzed from the perspectives of electron microscope (FE-SEM) (JSM-6700 F, JEOL Ltd, Japan). The
reception and conduction mechanisms. element mapping was performed by Transmission Electron Microscope
(TEM) (JEM-2010, JEOL Ltd, Japan). The surface composition was
2. Experimental details checked by X-ray Photoelectron Spectrometer (XPS) with mono­
chromated Al Kα radiation (Escalab 250, Thermo Electron Corporation,
2.1. Preparation of sensing materials USA). The BET surface area was obtained by Surface Area and Porosity
Analyzer (ASAP 2460, Micromeritics Instruments Corporation, USA).
The precursor of Zn-based MOF-5 powders was synthesized by a
solvothermal method. Specifically, 10 mmol Zn(NO3)2⋅6H2O and 2.3. Fabrication and measurements of sensors
5 mmol 1,4-benzenedicarboxylic acid (H2BDC) were added into the
mixed solvent of 50 mL dimethylformamide (DMF) and 10 mL methanol. Indirectly heated gas sensors were fabricated by coating the pristine
After ultrasonic treatment for 15 min., the obtained mixture in a 100 mL ZnO nanoparticles and ZnFe2O4/(Fe-ZnO) nanocomposites on ceramic
autoclave was kept in an oven at 150 ◦ C for 12 h. Upon centrifugation, tubes, followed by calcination in air at 200 ◦ C for 2 h and aging at 200 ◦ C
the obtained precipitate was washed by DMF for three times and dried at for 48 h to achieve stable sensor configuration. The details about the
160 ◦ C for 12 h to get the MOF-5 powders. The final products of pristine fabrication process and ceramic tube can be found in our previous work
ZnO nanoparticles were prepared by heating the as-prepared MOF-5 [14]. The acetone sensing measurement of the sensors was performed in
powders in air at 500 ◦ C for 2 h. Setting the mole ratio of Fe/Zn as x an atmosphere with the relative humidity of 15 % by an Intelligent Gas
(x = 0, 0.05, 0.10, 0.15, 0.20, 0.25 and 0.30), MOF-5 powders were Sensing Analysis System (CGS-8, Beijing Elite Tech Co., Ltd. China). The
immersed in aqueous solution of Fe(NO3)3⋅9H2O containing the gas response of the sensors was defined as the ratio of resistance in dry
required amount of Fe. The mixed solution was under constant stirring air (Ra) over that in target gas (Rg), and the response and recovery time
at 80 ◦ C in a water bath until it was totally dried. The obtained powers were defined as the time taken to achieve 90 % of the total resistance
were finely ground and calcined in air at 500 ◦ C for 2 h to get a series of variation in the case of adsorption and desorption, respectively.

2
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

Fig. 2. SEM images of (a): as-prepared MOF-5 powder; (b): typical MOF-5 cube, (c) and (d): typical ZnO microspheres and corresponding local magnified image; (e)
and (f): MOF-5 powder after immersion in Fe(NO3)3 aqueous solution with x = 0.15 and corresponding local magnified image.

3
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

Fig. 3. XRD patterns and TEM images of (a) and (c): ZnO nanoparticles; (b) and (d): ZnFe2O4/(Fe-ZnO) nanocomposite with x = 0.15.

3. Results and discussion Fig 0.2d indicates that the porous microsphere consists of pillars made of
ZnO nanoparticles. As for the sample after immersion in Fe(NO3)3
3.1. Characterization of sensing materials aqueous solution, it is composed of MIL-53(Fe) backbones covered by
the sheet bars of Zinc hydroxide nitrate hydrate, as revealed by Fig. 2e
Fig. 1a and 1b show the XRD patterns of the as-prepared MOF-5 and and 2f.
that after immersion in Fe(NO3)3 aqueous solution with x = 0.15, After calcination at 500 ◦ C for 2 h, the as-prepared MOF-5 powder
respectively. The as-prepared MOF-5 showed the main phase of MOF-5 turned into pristine ZnO nanoparticles and those after immersion in Fe
and the trace phase of hexaganol ZnO (PDF#89− 1397). The diffraction (NO3)3 aqueous solution became the ZnFe2O4/(Fe-ZnO) nano­
peaks at around 7.1◦ , 9.8◦ , 13.9◦ , and 15.6◦ correspond to the charac­ composites. Fig. 3 shows the XRD patterns and TEM images of the
teristic diffractions of (200), (220), (400) and (420) from MOF-5 pristine ZnO nanoparticles and the ZnFe2O4/(Fe-ZnO) nanocomposite
[15–17], while the peaks at 31.7◦ and 36.4◦ are assigned to the with x = 0.15. The phase of hexaganol ZnO (PDF#89− 1397) could be
featured (100) and (101) peaks of hexaganol ZnO [11–13]. The well observed in both samples. Based on the ZnO diffraction peaks, the lattice
resolved peaks imply the high crystallinity of MOF-5, and the stronger parameters, unit cell volume, strain and average crystallite size were
diffractions of (400) and (200) than other diffractions indicate the calculated using the Scherrer formula. Compared with the pristine ZnO,
preferred orientation along (100) for the microcrystalline MOF-5. As for the unit cell volume of ZnO in the ZnFe2O4/(Fe-ZnO) nanocomposite
the sample after immersion in Fe(NO3)3 aqueous solution, the mixed decreased from 47.80 Å3 to 47.75 Å3, the strain increased from 0.037 %
phases of MIL-53(Fe) [18–20] and Zinc hydroxide nitrate hydrate (Zn to 0.079 %, and the average crystallite size increased from 36.5 nm to
(OH)(NO3)(H2O)) appear instead of MOF-5 and ZnO. This can be 74.0 nm. These facts indicate that the smaller Fe ions might incorporate
explained by the decomposition of MOF-5 into Zinc salt and H2BDC in into the ZnO lattice and occupy some Zn2+ sites, which in turn resulted
aqueous solution, the concomitant reconstruction of MIL-53(Fe) by Fe3+ in the contraction of lattice and the increase of strain inside the ZnO
ions and H2BDC, and the transformation of ZnO into Zinc salt in Fe nanoparticles. Besides hexaganol ZnO, the nanocomposite showed the
(NO3)3 aqueous solution. independent phase of cubic ZnFe2O4 (PDF#74− 2397) [11–13]. As
The surface morphology of the as-prepared MOF-5 powder and that typically indicated by the green hexagons in the TEM images, the ZnO
after immersion in Fe(NO3)3 aqueous solution with x = 0.15 was nanoparticles in both samples mainly existed in the shape of hexagons
checked by FE-SEM, and the results are displayed in Fig. 2. As seen from which are in accordance with its crystalline structure. While the smaller
Fig. 2a, the as-prepared MOF-5 powder is mainly composed of MOF-5 nanoparticles inside the red circles and those dispersed in-between the
cubes, on the surface of which some ZnO microspheres grow. The ZnO nanoparticles are assigned to be the ZnFe2O4 nanoparticles. On the
typical MOF-5 cube and ZnO microsphere are presented in Fig. 2b and basis of 100 particles in clear outlines, the mean particle sizes of ZnO in
2c, respectively. The local magnified image of the ZnO microsphere in the pristine ZnO nanoparticles and ZnFe2O4/(Fe-ZnO) nanocomposite

4
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

Fig. 4. TEM (a) and corresponding SAED (b) images of the ZnFe2O4/(Fe-ZnO) nanocomposite with x = 0.15.

5
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

Fig. 5. Element mapping and STEM images for the ZnFe2O4/(Fe-ZnO) nanocomposite with x = 0.15 (BF, Bright Field; HAADF, High Angle Annular Dark Field).

Fig. 6. Whole survey, Zn2p, O1s, and Fe2p core level XPS of the pristine ZnO nanoparticles and ZnFe2O4/(Fe-ZnO) nanocomposite with x = 0.15 (oct, octahedral; tet,
tetrahedral; sat, satellite).

6
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

Fig. 7. SEM images of ZnO nanoparticles (a) and (b): in the microstructure of cube and corresponding local magnified image; (c) and (d): in the microstructure of
sphere and corresponding local magnified image; (e)-(j): ZnFe2O4/(Fe-ZnO) nanocomposites with x = 0.05, 0.10, 0.15, 0.20, 0.25 and 0.30, respectively.

7
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

and sphere are in the porous structure and compose of ZnO nano­
particles in uniformly distributed size. Fig. 7e-7 j display the SEM images
of the ZnFe2O4/(Fe-ZnO) nanocomposites with x = 0.05, 0.10, 0.15,
0.20, 0.25 and 0.30, respectively, which all showed the porous structure
of nanoparticles. Inferred by the analyses above, the bigger particles in
these images should be ZnO nanoparticles, while the smaller ones should
be ZnFe2O4 nanoparticles. To further check the effect of Fe/Zn mole
ratios on the microstructure of nanocomposites, BET measurement was
performed, and the results of BET surface area and mean pore size for
each sample are depicted in Fig. 8. The pristine ZnO nanoparticles
possess the largest surface area and smallest pore size due to its pre­
served microstructure of MOF-5 cubes and ZnO spheres. As for the
ZnFe2O4/(Fe-ZnO) nanocomposites, with the increase of Fe/Zn mole
ratio, the BET surface area increases at first, reaches a maximum at
x = 0.15, then decreases again. Basically, a larger surface area corre­
Fig. 8. BET surface area (left) and average pore size (right) of ZnFe2O4/(Fe- sponds to a smaller mean pore size.
ZnO) nanocomposites with different x values.
3.2. Acetone sensing performance and mechanism
and the mean particle size of ZnFe2O4 in the nanocomposite were esti­
mated to be 43.5 nm, 63.6 nm and 15.5 nm, respectively, which were Fig. 9 shows the acetone sensing characteristics of the pristine ZnO
close to the average crystallite sizes of 36.5 nm, 74.0 nm and 18.5 nm nanoparticles and the ZnFe2O4/(Fe-ZnO) nanocomposite-based sensors.
estimated by the XRD. The polycrystalline nature of the Fe-ZnO and The temperature dependence of resistance in Fig. 9a shows that all the
ZnFe2O4 nanoparticles was further verified by their characteristic nanocomposites showed higher resistance than the pristine ZnO nano­
diffraction rings in the SAED image in Fig. 4. particles due to the existence of ZnFe2O4 insulating phase. For the same
Since the diffraction peaks of Fe2.93O4 and ZnFe2O4 are very close to reason, higher ratios of Fe/Zn in the nanocomposites resulted in higher
each other, it is difficult to distinguish merely from the XRD patterns. resistance. From the operating temperature dependence of the gas
Therefore, the element mapping in combination with the STEM in the response to 100 ppm acetone in Fig. 9b, the prime working temperature
Bright Field (BF) and High Angle Annular Dark Field (HAADF) were (TP) where the gas response reaches its maximum was determined to be
performed on the nanocomposite with x = 0.15, and the results were 270 ◦ C for the pristine ZnO sensor, while the TP shifted to lower tem­
displayed in Fig. 5. Apparently, Zn, Fe and O could be observed in the perature with the increase of x value for the ZnFe2O4/(Fe-ZnO) nano­
yellow circle, which means that the inside particle should be ZnFe2O4 composites. The highest gas response of 30.8, which was almost four
rather than Fe2.93O4. The particles with Fe atoms in the dense state times the value of 8.2 for the ZnO sensor at 270 ◦ C, was achieved by the
should be ZnFe2O4 as well, while the particles having dense Zn and O nanocomposite with x = 0.15 at the TP of 190 ◦ C. The variation in gas
atoms as well as sparsely distributed Fe atoms should be the Fe-doped response over the variation in concentration of target gas reflects the
ZnO nanoparticles. sensitivity of the sensor. It can be seen from Fig. 9c that at the TP of each
Fig. 6 shows the whole survey and elemental XPS in the core level of sensor the nanocomposite with x = 0.15 exhibited the highest sensitivity
the pristine ZnO nanoparticles and the ZnFe2O4/(Fe-ZnO) nano­ among all sensors in the range of 10− 200 ppm. For practical application,
composite with x = 0.15, with the binding energy (BE) in all spectra the selectivity of the sensors must be considered. Fig. 9d displays the gas
charge-shifted relative to the C–C bond at 284.8 eV. Except the carbon responses to 100 ppm different target gases at the TP of each sensor. The
from the unavoidable contamination, all peaks in the whole survey ratio of the gas response to acetone over the response to EG, DMF,
spectra (Fig. 6a) can be assigned to the Zn and O core level peaks and ammonia water, N-hexane or carbon dioxide was much higher for the
Auger peaks, confirming the purity of elements in both samples. In ZnFe2O4/(Fe-ZnO) sensors than that for the pristine ZnO sensor, which
Fig. 6b, the peaks at the BE of 1021.3 eV and 1044.4 eV in both samples means that the ZnFe2O4/(Fe-ZnO) nanocomposites showed higher
are assigned to the Zn2p3/2 and 2p1/2 spin-orbit components in ZnO, selectivity towards acetone than that of ZnO nanoparticles. The reason
respectively, while the peaks at the BE of 1021.9 eV and 1045.4 eV for for the enhanced selectivity towards target gas is still unclear now,
the ZnFe2O4/(Fe-ZnO) nanocomposite were ascribed to the Zn2p3/2 and which is supposed to be associated with some intrinsic properties of
2p1/2 spin-orbit components in ZnFe2O4 [2,12]. In Fig. 6c, the peak at target gases, such as the reaction activity in terms of bond energy [23,
the BE of 530.1 eV is assigned to the lattice oxygen (OL) in ZnO or 24], the dipole moment in the molecular structure [25], the molecular
ZnFe2O4, while the peak around 531.6 eV is ascribed to the chemisorbed size and molecular weight [26]. Besides, the porous structure of the
oxygen species (OC) on the surface [10,21]. Because the Fe in ZnFe2O4 or sensing matrix in terms of pore size is reported to determine the diffu­
ZnO can result in the stronger Fe-O bond than Zn-O bond on the surface, sion type of target gas by comparing the mean free path of target gas and
the BE of OC is increased to 531.7 eV in the composite. The Fe2p core mean pore size, which will also contribute to the selectivity of sensing
level XPS for the ZnFe2O4/(Fe-ZnO) nanocomposite is displayed in material [27].
Fig. 6d, which shows Fe in three different valence states [22]. The For the pristine ZnO and the ZnFe2O4/(Fe-ZnO) nanocomposite with
normal Fe3+ and reduced Fe2+ species should originate from the Fe ions x = 0.15, Fig. 9e depicts the dynamic responses of resistance to 100 ppm
in the ZnFe2O4 and ZnO lattices, while the metallic Fe atoms might be acetone at each TP in three consecutive measurements. Both sensors
sparsely located on the surface of ZnO nanoparticles. The Fe species in showed good recovery, stability, and reproducibility, and the ZnFe2O4/
the reduced state should be caused by the reducing carbide and nitride (Fe-ZnO) sensor displayed even more stable resistance than the ZnO
emitted during the annealing process. sensor after gas was taken out. The mean response time for the ZnFe2O4/
The surface morphology of the pristine ZnO and the ZnFe2O4/(Fe- (Fe-ZnO) sensor was calculated to be 4.7 s which was longer than the
ZnO) nanocomposites with x = 0.05, 0.10, 0.15, 0.20, 0.25 and 0.30 3.3 s for the ZnO sensor, while the mean recovery time of 10.3 s was
were displayed by the FE-SEM images in Fig. 7. As shown in Fig. 7a and smaller than the 11.0 s for the ZnO sensor. The above sensing mea­
7c, the pristine ZnO nanoparticles preserved the microstructure of cube surements were all performed in an atmosphere of 15 % relative hu­
and sphere as in the as-prepared MOF-5 powder before calcination. midity. To study the effect of relative humidity (RH) on the sensing
While the local magnified images in Fig. 7b and 7d indicate that the cube performance, the RH dependence of gas responses to 100 ppm acetone
for the pristine ZnO and the nanocomposite with x = 0.15 was examined

8
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

Fig. 9. For the ZnFe2O4/(Fe-ZnO) nanocomposites-based sensors with x = 0, 0.05, 0.10, 0.15, 0.20, 0.25 and 0.30, (a): temperature dependence of resistance from
170 to 290 ◦ C; (b): temperature dependence of gas response to 100 ppm acetone from 170 to 290 ◦ C; (c): gas responses to different concentration of acetone at each
prime working temperature; (d): gas responses to 100 ppm different target gases at each prime working temperature. (e): dynamic responses of resistance to 100 ppm
acetone at each prime working temperature in three consecutive measurements for x = 0 and 0.15; (f): relative humidity dependence of gas responses to 100 ppm
acetone at 190 ◦ C for x = 0 and 0.15.

9
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

Table 1
Lattice parameters, unit cell volume (V), average crystallite size (DC), strain (σ), mean particle size (DP) of ZnO in the pristine ZnO nanopartilces (ZnOP) and ZnFe2O4/
(Fe-ZnO) nanocomposite (ZnOC) with x = 0.15, and these parameters of ZnFe2O4 in the nanocomposite.
Phase a (Å) b (Å) c (Å) V (Å3) DC (nm) σ (%) DP (nm)

ZnOP 3.2530 3.2530 5.2127 47.77 36.5 ± 0.6 0.037 43.5


ZnOC 3.2529 3.2529 5.2103 47.75 74.0 ± 5.5 0.079 63.6
ZnFe2O4 8.4716 8.4716 8.4716 607.99 18.5 ± 1.3 0 15.5

Table 2
Comparison between the acetone sensing performance of ZnO-based composites reported in literature, including prime working temperature (TP), gas response, and
response/recovery time (TRes/Trec).
Sensing material Preparation method Structure TP Response TRes/Trec Ref.

CeO2/ZnO Solvothermal Dual porous composite 320℃ 16.0− 5 ppm / [29]


α-Fe2O3/ZnO Hydrothermal and sol-gel Nanotubes 369℃ 32.0− 100 ppm / [30]
α-Fe2O3/ZnO Hydrothermal Core-shell nanospindles 280℃ 12.0− 100 ppm / [31]
NiO/ZnO Solution-based Flower-like 300℃ 23.5− 100 ppm / [32]
NiO/ZnO Chemical bath deposition Flower-like 330℃ 13.5− 100 ppm* / [33]
NiO/ZnO Solvothermal Yolk-double-shelled microspheres 260℃ 7.5− 200 ppm 8s/13 s [34]
ZnCo2O4/ZnO Solution-based ZnCo2O4 nanodots on ZnO nanorods 240℃ 15.3− 10ppm 7.7 s/10.3 s [35]
SmFeO3/ZnO Solution-based Nanocomposite 350℃ 45− 10ppm 10− 15 s /250− 400 s [36]
ZnFe2O4/ZnO Prussian Blue analogue derived Hollow microspheres 140℃ 14.0− 100 ppm 5.2 s/12.8s [2]
ZnFe2O4/ZnO Thermal annealing of Fe(III)MOF-5 Hollow nanocages 290℃ 25.8− 100 ppm 8s/32 s [11]
ZnFe2O4/ZnO Solution-based ZnFe2O4 nanoparticles on ZnO microflowers 250℃ 8.3− 50ppm 2 s/25s [13]
ZnFe2O4/ZnO Solvothermal Core-shell hollow microspheres 280℃ 18.0− 80ppm* 8s/30 s [37]
ZnFe2O4/ZnO Hydrolyzation of MOF-5 Nanocomposite 190 ◦ C 30.8− 100 ppm 4.7 s/10.3 s This work
*
Inferred from data in graphic.

at the same temperature of 190 ◦ C to minimize the effect of temperature


C3 H6 O + 8On−ads →3CO2 + 3H2 O + 8ne− (5)
on the adsorbing state of water molecules on both samples. The results in
Fig. 9f indicate that the RH had a considerable negative effect on the gas The improved acetone sensing performance of the ZnFe2O4/(Fe-
response of both samples owing to the competing relationship between ZnO) nanocomposite can be analyzed by the reception and transduction
the water and oxygen molecules on the surface adsorbing sites as well as mechanisms. Firstly, the mismatched lattice caused by the incorporation
the reverse contribution to the width of electron depletion layer [28]. of Fe into the ZnO lattice, the interfaces between ZnFe2O4 and Fe-ZnO
Table 1 nanoparticles, and the grain boundaries in-between ZnFe2O4 nano­
The acetone sensing characteristics of ZnO-based composites re­ particles are all high energy sites favoring the adsorption of oxygen and
ported in literature are compared in Table 2. Compared with the CeO2/ acetone. Therefore, the ZnFe2O4/(Fe-ZnO) nanocomposite with x = 0.15
ZnO, α-Fe2O3/ZnO, NiO/ZnO, ZnCo2O4/ZnO, SmFeO3/ZnO composites which has the largest BET surface area exhibited the highest gas
[29–36], our ZnFe2O4/ZnO nanocomposite exhibited much lower TP response due to its abundance of adsorption sites and higher surface
and faster dynamic response with relatively high gas response. As for oxygen coverage. Secondly, ZnFe2O4 is an n-type semiconductor with
other ZnFe2O4/ZnO composites reported in literature [2,11,13,37], our higher Fermi level than that of ZnO, thus electrons would flow from
sample displayed the highest gas response with comparatively lower TP ZnFe2O4 to ZnO across the interface of ZnFe2O4/(Fe-ZnO) hetero­
and faster dynamic response. Therefore, the ZnFe2O4/ZnO nano­ structure. This charge redistribution results in electron accumulation in
composite in our work, as sensing material, is featured by high response, the ZnO and electron depletion in the ZnFe2O4. More electrons on the
fast response/recovery speed at low prime working temperature with ZnO surface would benefit the oxygen adsorption from air and result in
good reliability. wider depletion layer. Given the small particle size of 15.5 nm, the
When the n-type ZnFe2O4/(Fe-ZnO) nanocomposite is exposed in air, remained electrons in the ZnFe2O4 nanoparticles could be completely
the oxygen in air will adsorb on the surface of ZnFe2O4 and Fe-ZnO depleted by the adsorbed oxygen [39]. The wider depletion layer at the
nanoparticles and undergo the following reactions as temperature rises. ZnO surface, the completely depleted ZnFe2O4, as well as the higher and
wider potential barrier at the interfaces give rise to a larger measurable
O2(gas) ↔ O2(ads) (1)
variation in resistance when the adsorbed oxygen species are removed
O2(ads) + e− ↔ O−2(ads) (2) by surface interactions with acetone.

O−2(ads) + e− ↔ 2O−(ads) (3) 4. Conclusion

O−(ads) + e− ↔ O2−(ads) (4) A series of ZnFe2O4/(Fe-ZnO) nanocomposites (the mole ratio of Fe/
Zn = 0, 0.05, 0.10, 0.15, 0.20, 0.25 and 0.30) was prepared by a
As illustrated in Fig. 10a, a depletion layer would be formed on the solution-based method via the hydrolyzation of Zn-based MOF-5 in Fe
surface of ZnFe2O4 and Fe-ZnO nanoparticles due to the oxygen (NO3)3 aqueous solution. XRD, TEM, SAED, Element mapping and XPS
adsorption. Given the weak nature of direct interaction between the confirmed the composition of cubic ZnFe2O4 and Fe-doped hexagonal
acetone and surface, the conductivity is considered to be mainly ZnO in the nanocomposites. Compared with the MOF-5-derived pristine
controlled by the indirect chemical reaction between adsorbed oxygen ZnO nanoparticles, the ZnFe2O4/(Fe-ZnO) nanocomposites, especially
species (On−
ads ) and introduced acetone as Eq. (5) [38], which will lead to the one with Fe/La = 0.15, exhibited enhanced gas response, higher
the decrease in the width of electron depletion layer for the ZnFe2O4 and sensitivity, improved stability and reproducibility, faster recovery, and
Fe-ZnO nanoparticles, as illustrated in Fig. 10b. better selectivity to acetone at lower prime working temperature. From
the perspectives of reception and transduction mechanisms, the

10
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

Fig. 10. Schematic diagram of structure and band for the ZnFe2O4/(Fe-ZnO) nanocomposite-based sensor in air (a) and in acetone (b).

11
E. Cao et al. Sensors and Actuators: B. Chemical 325 (2020) 128783

enhanced acetone sensing performance should be contributed by the [13] C. Liu, B. Wang, T. Wang, J. Liu, P. Sun, X. Chuai, G. Lu, Enhanced gas sensing
characteristics of the flower-like ZnFe2O4/ZnO microstructures, Sens. Actuators B
synergistic effect of Fe-doping and heterostructure formation on the
Chem. 248 (2017) 902–909.
material’s porous structure, adsorption sites, and modulation of poten­ [14] K. Cao, E. Cao, Y. Zhang, W. Hao, L. Sun, H. Peng, The influence of
tial barrier at the interfaces. Looking forward, the preparation of nonstoichiometry on electrical transport and ethanol sensing characteristics for
nanocomposite through the hydrolyzation of MOF material in transition nanocrystalline LaFexO3− δ sensors, Sens. Actuators B Chem. 230 (2016) 592–599.
[15] J. Hafizovic, M. Bjørgen, U. Olsbye, P.D.C. Dietzel, S. Bordiga, C. Prestipino,
metal aqueous solution may be a facile and effective way in improving C. Lamberti, K.P. Lillerud, The inconsistency in adsorption properties and powder
the gas sensing performance of MOF-derived nanoparticles by the syn­ xrd data of mof-5 is rationalized by framework interpenetration and the presence
ergistic effect of doping and heterostructure formation, while more work of organic and inorganic species in the nanocavities, J. Am. Chem. Soc. 129 (2007)
3612–3620.
needs to be done to weaken the strong dependence of gas response on [16] Z. Zhao, Z. Li, Y.S. Lin, Adsorption and diffusion of carbon dioxide on metal-
the surrounding humidity for practical application. organic framework (mof-5), Ind. Eng. Chem. Res. 48 (2009) 10015–10020.
[17] Y. Wu, H. Pang, W. Yao, X. Wang, S. Yu, Z. Yu, X. Wang, Synthesis of rod-like
metal-organic framework (MOF-5) nanomaterial for efficient removal of U(VI):
CRediT authorship contribution statement batch experiments and spectroscopy study, Sci. Bull. 63 (2018) 831–839.
[18] Y. Zhang, J. Zhou, J. Chen, X. Feng, W. Cai, Rapid degradation of tetracycline
Ensi Cao: Conceptualization, Methodology, Writing - original draft, hydrochloride by heterogeneous photocatalysis coupling persulfate oxidation with
MIL-53(Fe) under visible light irradiation, J. Hazard. Mater. 392 (2020), 122315.
Writing - review & editing. Zhaoqing Guo: Investigation. Guoqing [19] G. Chaturvedi, A. Kaur, S.K. Kansal, CdS-decorated MIL-53(fe) microrods with
Song: Validation. Yongjia Zhang: Data curation. Wentao Hao: Re­ enhanced visible light photocatalytic performance for the degradation of ketorolac
sources. Li Sun: Project administration. Zhongquan Nie: Formal tromethamine and mechanism insight, J. Phys. Chem. C 123 (2019) 16857–16867.
[20] F.L. Li, Q. Shao, X. Huang, J.P. Lang, Nanoscale trimetallic metal-organic
analysis.
frameworks enable efficient oxygen evolution electrocatalysis, Angew. Chem. Int.
Ed. Engl. 57 (2018) 1888–1892.
Declaration of Competing Interest [21] X. Xing, Y. Yang, Z. Yan, Y. Hu, T. Zou, Z. Wang, Y. Wang, CdO-Ag-ZnO
nanocomposites with hierarchically porous structure for effective VOCs gas-sensing
properties, Ceram. Int. 45 (2019) 4322–4334.
The authors declare that they have no known competing financial [22] P. Tan, Active phase, catalytic activity, and induction period of Fe/zeolite material
interests or personal relationships that could have appeared to influence in nonoxidative aromatization of methane, J. Catal. 338 (2016) 21–29.
the work reported in this paper. [23] Q. Xu, D. Ju, Z. Zhang, S. Yuan, J. Zhang, H. Xu, B. Cao, Near room-temperature
triethylamine sensor constructed with CuO/ZnO P-N heterostructural nanorods
directly on flat electrode, Sens. Actuators B Chem. 225 (2016) 16–23.
Acknowledgments [24] X. Song, Q. Xu, H. Xu, B. Cao, Highly sensitive gold-decorated zinc oxide nanorods
sensor for triethylamine working at near room temperature, J. Colloid Interface
Sci. 499 (2017) 67–75.
This work was supported by the Natural Science Foundation of [25] R. Yoo, A.T. Güntner, Y. Park, H.J. Rim, H.-S. Lee, W. Lee, Sensing of acetone by Al-
Shanxi Province (201901D111117 and 201901D111126), and National doped ZnO, Sens. Actuators B Chem. 283 (2019) 107–115.
Natural Science Foundation of China (11604236 and 11974258). [26] W. Li, Z. Feng, E. Dai, J. Xu, G. Bai, Organic vapour sensing properties of area-
ordered and size-controlled silicon nanopillar, Sensors 16 (2016) 1880–1887.
[27] E. Cao, Z. Chu, H. Wang, W. Hao, L. Sun, Y. Zhang, Effect of film thickness on the
References electrical and ethanol sensing characteristics of LaFeO3 nanoparticle-based thick
film sensors, Ceram. Int. 44 (2018) 7180–7185.
[1] Acetone: Health Information Summary, New Hampshire, Department of [28] G. Heiland, D. Kohl, in: T. Seiyama (Ed.), (Ed.), Chemical Sensor Technology, vol.
Environmental Service, 2005. 1, Kodansha, Tokyo, 1988, pp. 15–38. Chapter 2).
[2] X.-Z. Song, L. Qiao, K.-M. Sun, Z. Tan, W. Ma, X.-L. Kang, F.-F. Sun, T. Huang, X.- [29] Q. Meng, J. Cui, Y. Tang, Z. Han, K. Zhao, G. Zhang, Q. Diao, Solvothermal
F. Wang, Triple-shelled ZnO/ZnFe2O4 heterojunctional hollow microspheres synthesis of dual-porous CeO2-ZnO composite and its enhanced acetone sensing
derived from Prussian Blue analogue as high-performance acetone sensors, Sens. performance, Ceram. Int. 45 (2019) 4103–4107.
Actuators B Chem. 256 (2018) 374–382. [30] F. Wang, J. Liu, X. Wang, J. Kong, S. Qiu, G. Lu, C. He, Alpha-Fe2O3@ZnO
[3] Q. Ma, S. Chu, Y. Liu, Y. Chen, J. Song, H. Li, J. Wang, Q. Che, G. Wang, Y. Fang, heterostructured nanotubes for gas sensing, Mater. Lett. 76 (2012) 159–161.
Construction of SnS2/ZIF-8 derived flower-like porous SnO2/ZnO heterostructures with [31] J. Zhang, X. Liu, L. Wang, T. Yang, X. Guo, S. Wu, S. Wang, S. Zhang, Synthesis and
enhanced triethylamine gas sensing performance, Mater. Lett. 236 (2019) 452–455. gas sensing properties of α-Fe2O3@ZnO core-shell nanospindles, Nanotechnology
[4] P. Wang, T. Dong, C. Jia, P. Yang, Ultraselective acetone-gas sensor based ZnO 22 (2011), 185501.
flowers functionalized by Au nanoparticle loading on certain facet, Sens. Actuators [32] C. Liu, B. Wang, T. Liu, P. Sun, Y. Gao, F. Liu, G. Lu, Facile synthesis and gas
B Chem. 288 (2019) 1–11. sensing properties of the flower-like NiO-decorated ZnO microstructures, Sens.
[5] Y. Li, S. Wang, P. Hao, J. Tian, H. Cui, X. Wang, Soft-templated formation of double- Actuators B Chem. 235 (2016) 294–301.
shelled ZnO hollow microspheres for acetone gas sensing at low concentration/near [33] Y. Liu, G. Li, R. Mi, C. Deng, P. Gao, An environment-benign method for the
room temperature, Sens. Actuators B Chem. 273 (2018) 751–759. synthesis of p-NiO/n-ZnO heterostructure with excellent performance for gas
[6] X. Xie, X. Wang, J. Tian, X. Song, N. Wei, H. Cui, Growth of porous ZnO single sensing and photocatalysis, Sens. Actuators B Chem. 191 (2014) 537–544.
crystal hierarchical architectures with ultrahigh sensing performances to ethanol [34] R. Zhang, J. Shi, T. Zhou, J. Tu, T. Zhang, A yolk-double-shelled heterostructure-
and acetone gases, Ceram. Int. 43 (2017) 1121–1128. based sensor for acetone detecting application, J. Colloid Interface Sci. 539 (2019)
[7] X.-F. Wang, X.-Z. Song, K.-M. Sun, L. Cheng, W. Ma, MOFs-derived porous 490–496.
nanomaterials for gas sensing, Polyhedron 152 (2018) 155–163. [35] H. Qin, T. Liu, J. Liu, Q. Liu, R. Li, H. Zhang, J. Wang, Fabrication of uniform 1-D
[8] W. Li, X. Wu, N. Han, J. Chen, X. Qian, Y. Deng, W. Tang, Y. Chen, MOF-derived ZnO/ZnCo2O4 nano-composite and enhanced properties in gas sensing detection,
hierarchical hollow ZnO nanocages with enhanced low-concentration VOCs gas- Mater. Chem. Phys. 228 (2019) 66–74.
sensing performance, Sens. Actuators B Chem. 225 (2016) 158–166. [36] Z. Anajafi, M. Naseri, G. Neri, Acetone sensing behavior of p-SmFeO3/n-ZnO
[9] W. Li, X. Wu, N. Han, J. Chen, W. Tang, Y. Chen, Core-shell Au@ZnO nanoparticles nanocomposite synthesized by thermal treatment method, Sens. Actuators B Chem.
derived from Au@MOF and their sub-ppm level acetone gas-sensing performance, 304 (2020), 127252.
Powder Technol. 304 (2016) 241–247. [37] Y. Hu, H. Wang, D. Liu, G. Lin, J. Wan, H. Jiang, X. Lai, S. Hao, X. Liu, Lychee-like
[10] H. Fu, X. Wang, P. Wang, Z. Wang, H. Ren, C.C. Wang, Enhanced acetone sensing ZnO/ZnFe2O4 core-shell hollow microsphere for improving acetone gas sensing
performance of Au nanoparticle modified porous tube-like ZnO derived from rod- performance, Ceram. Int. 46 (2020) 5960–5967.
like ZIF-L, Dalton Trans. 47 (2018) 9014–9020. [38] J.M. Walker, S.A. Akbar, P.A. Morris, Synergistic effects in gas sensing
[11] X. Wang, S. Zhang, M. Shao, J. Huang, X. Deng, P. Hou, X. Xu, Fabrication of ZnO/ semiconducting oxide nano-heterostructures: a review, Sens. Actuators B Chem.
ZnFe2O4 hollow nanocages through metal organic frameworks route with 286 (2019) 624–640.
enhanced gas sensing properties, Sens. Actuators B Chem. 251 (2017) 27–33. [39] D.R. Miller, S.A. Akbar, P.A. Morris, Nanoscale metal oxide-based heterojunctions
[12] X. Li, C. Han, D. Lu, C. Shao, X. Li, Y. Liu, Highly electron-depleted ZnO/ZnFe2O4/ for gas sensing: a review, Sens. Actuators B Chem. 204 (2014) 250–272.
Au hollow meshes as an advanced material for gas sensing application, Sens.
Actuators B Chem. 297 (2019), 126769.

12

You might also like