You are on page 1of 28

Journal Pre-proof

Urea mediated synthesis and acetone-sensing properties of ultrathin


porous ZnO nanoplates

Lai Van Duy, Nguyen Van Duy, Chu Manh Hung, Nguyen Duc Hoa,
Nguyen Quang Dich

PII: S2352-4928(20)32456-9
DOI: https://doi.org/10.1016/j.mtcomm.2020.101445
Reference: MTCOMM 101445

To appear in: Materials Today Communications

Received Date: 24 May 2020


Revised Date: 8 June 2020
Accepted Date: 7 July 2020

Please cite this article as: Van Duy L, Van Duy N, Hung CM, Hoa ND, Dich NQ, Urea
mediated synthesis and acetone-sensing properties of ultrathin porous ZnO nanoplates,
Materials Today Communications (2020), doi: https://doi.org/10.1016/j.mtcomm.2020.101445

This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2020 Published by Elsevier.


Urea mediated synthesis and acetone-sensing properties of ultrathin porous

ZnO nanoplates

Lai Van Duy1, Nguyen Van Duy1*, Chu Manh Hung1, Nguyen Duc Hoa1,

Nguyen Quang Dich2,*

1
International Training Institute for Materials Science (ITIMS), Hanoi University of Science and Technology,

of
No. 1, Dai Co Viet, Hai Ba Trung, Hanoi, Vietnam;

2
Institute for Control Engineering and Automation, Hanoi University of Science and Technology (HUST), 1-

ro
Daicoviet, Hanoi, Vietnam

Corresponding author: ndhoa@itims.edu.vn / dich.nguyenquang@hust.edu.vn


-p
re
Graphical abstract:

Ultrathin porous ZnO nanoplates processed excellent acetone sensing performance

for breath analysis towards diabetes diagnose


lP
na
ur

Highlights
Jo

 Ultrathin porous ZnO nanoplates were synthesized for acetone gas sensor
 The ZnO nanoplate has a pore size of 2.5 nm and a thickness of 15 nm
 The sensor showed excellent acetone sensing performance with detection limit
of 45 ppb
 The sensor is suitable for diabetes diagnosing via breath analysis

Abstract

Facile synthesis of advanced porous nanomaterials for enhanced gas sensing performance has gained increasing

interest. Here, we report our study on the preparation of ultrathin porous ZnO nanoplates by a urea mediated

hydrothermal method for acetone gas-sensor application. The synthesized ZnO nanomaterials were nanoplates

with a porous structure, high crystallinity, and ultrathin thickness of approximately 15 nm. The gas-sensing

characteristics of the ultrathin porous ZnO nanoplates were studied at 350–450 °C. The sensor showed the highest

of
response to acetone among the tested gases (methanol, ethanol, acetone, toluene and ammonia); the response to

2.5 ppm acetone was 1.8. The sensor also showed a low 45 ppb detection limit with excellent stability and

ro
selectivity at 450 °C. The acetone-sensing mechanism of the ultrathin porous ZnO nanoplates was also discussed.

Overall, the findings indicated that the ultrathin porous ZnO nanoplates were excellent materials for sensing

-p
applications (e.g., breath analysis) of volatile organic compounds.

Keywords: Ultrathin porous ZnO, nanoplates, VOCs, Acetone, Gas sensor.


re
1. Introduction
lP

Exhaled breath contains around 500 various volatile organic compounds (VOCs) [1,2] but their

concentrations are different depending on the human health condition [3]. The sources of VOCs in exhaled breath
na

can be endogenous or exogenous [4], such as a metabolic deficiency, a pathological disorder [5], cellular

destruction or human body inflammation [6]. Monitoring the VOCs in exhaled breath is an instant method for non-

invasive medical diagnoses [3,6] because the presence of some VOCs can be considered biomarkers for a disease
ur

[7]. It was reported that differences in the concentration of exhaled VOCs can serve as a fingerprint for specific

diseases [4], including diabetes and kidney failure [3]. Recently, acetone has been used as an effective biomarker
Jo

in diagnosing diabetes [8] because its concentration in the exhaled breath of a diabetic patient is higher than that

in healthy people. The detection limit of an acetone sensor for exhaled breath analysis is 1.8 ppm [9]. Therefore,

many attempts have been made to develop a highly sensitive acetone gas sensor [10].
Resistive gas sensors present possibilities for disease diagnosis because they have some advantages, such as

small size, low cost, portability, robustness, high responses and real-time monitoring [4,11]. Different metal oxides

[12] were prepared for VOC gas sensor applications [13–17]. Zinc oxide (ZnO) [18], an n-type semiconductor has

been widely used as a gas sensing material thanks to its simple synthesis and high response to various gases [19].

Attempts have focused on synthesising different ZnO nanostructures for VOC sensors [8,19–27]. Nanomaterials

with a porous structure were prepared for gas-sensing applications [28] because they can promote the diffusion

path to enhance the adsorbed quantity of target gas molecules [29]. The porous structure also possesses high

activity in surface chemical reactions and, thus, provides high gas sensitivity [30]. Therefore, gas sensors fabricated

with a porous morphology are expected to exhibit high response with reduced response and recovery times [31].

Two-dimensional (2D) ZnO nanomaterials have also proven to have a good sensing performance because of their

of
increased proportion of exposed active planes [32,33]. In addition, decorating the surface of sensing materials with

ro
noble metal particles could also improve the sensitivity of ZnO nanostructures [34]. The ZnO nanostructures can

be easily prepared by co-precipitation, sol-gel, electrospinning and hydrothermal methods [35,36]. For example,

-p
the 2D ZnO nanostructures could be prepared using a hydrothermal method with low-cost equipment compared to

thermal evaporation [37]. ZnO nanosheets doped with Sn were prepared by a hydrothermal method [38], where
re
the synthesized ZnO nanosheets have a dense structure with a 20 nm thickness. The response (Ra/Rg) value to 200

ppm acetone at 320 °C was approximately 5.6. Recently, porous ZnO nanoplates with large pores were prepared
lP

for ethanol detection [19]. However, the ZnO plates were large, i.e., approximately 5 µm in size, with a thickness

of approximately 100 nm, leading to a low sensor response. Therefore, an easy, less toxic, scalable synthesis

method for preparing ultrathin porous ZnO nanoplates remains challenging. In addition, the acetone sensing
na

properties of porous ZnO nanosheets are rarely reported.

Here, we report our study on the urea mediated synthesis of ultrathin porous ZnO nanoplates for acetone gas
ur

sensor application. The ultrathin porous ZnO nanoplates were synthesized by a facile hydrothermal method and

subsequent annealing process. Urea was used as a mediating agent to control the pH and porosity and the less toxic
Jo

bioreagent zinc sulfate was used as a Zn2+ precursor. Ultrathin porous ZnO nanoplates exhibited superior

sensitivity to acetone with excellent stability and are suitable for application in diabetes diagnosis.
2. Experimental

The reagents used in this study included zinc sulfate heptahydrate (ZnSO4·7H2O), urea (CH4N2O) and

deionized water. The analytical grade reagents were used as received. The hydrothermal method was used to

prepare ultrathin porous ZnO nanoplates [33,34]. Briefly, a mixture of 0.28 g ZnSO4·7H2O and 1.2 g CH4N2O was

dissolved in 50 mL deionized water at room temperature by magnetic steering. The pH of the solution was adjusted

to 5 by further stirring for 15 min. This solution was poured into a 100 mL Teflon-lined stainless-steel autoclave

[19]. After capping, the autoclave was loaded in an electric oven for hydrothermal material growth at 220 °C for

24 h. The electric oven was then turned off and left to cool to room temperature naturally. The white precipitate

was washed with a mixture of deionized water and ethanol to remove the unreacted reagents. The hydrothermal

of
product was centrifuged at 4000 rpm for 20 min to collect the precipitate. After drying at 60 °C for 24 h, the final

product was annealed for 2 h in air at 600 °C. The synthesized materials were characterized by some advanced

ro
techniques [39], such as SEM & EDS (JEOL 7600F), XRD (Advance D8, Bruker) and HRTEM (JEM2100).

Optical properties were studied by Photoluminescence (PL) and Raman spectroscopy measurements.
-p
The sensors were prepared using a thick film technique, as reported in ref. [39]. Briefly, the ZnO nanoplates

were dispersed in N-vinylpyrrolidone solution using an ultrasonic bath. The colloidal solution was then cast onto
re
Pt-interdigitated electrodes to form the sensors. The gas-sensing properties were measured at different

temperatures using a home-made system [40]. During gas sensing measurement, the resistance of the sensor was
lP

measured while the tested gas was introduced. The sensor response was defined as S=Ra/Rg, where Ra and Rg are

the resistances of the sensor in air and analytic gas, respectively.


na

3. Results and discussion

The morphologies of the product observed by SEM images (Figure 1[A–B]) display the ultrathin porous
ur

structure of the ZnO nanoplates. Numerous randomly oriented nanoplates, micrometres in width and length, were

found. The top surface and the edges of the ZnO nanoplates can be seen in Figure 1[A]. However, the nanoplates
Jo

were very thin (Figure 1[B]), approximately 15 nm. We could not see the porous structure of the nanoplates

because of the low-resolution SEM image. The thickness of the porous ZnO nanoplates (15 nm) prepared in our

study was far less than that of the porous ZnO nanoflakes (~70 nm) [30]. This is an advantage because it is

comparable with the Debye length (~20 nm) of ZnO, thus it was expected to show high gas sensitivity [41]. The

mechanism for the formation of the porous ZnO nanoplates can be explained as follows. When dissolved in water,
the zinc sulphate heptahydrate formed the Zn2+ ions. During hydrothermal treatment at high temperature, the

CH4N2O decomposed into NH4OH and HNCO. Thus, the (OH)− group reacted with Zn2+ ions to form Zn(OH)2

flakes. The flakes grew with hydrothermal treatment and then decomposed into ZnO and generated the ultrathin

porous nanosheet structure. The crystal structure of the ZnO nanoplates was investigated by XRD and is shown in

Figure 1(C). The XRD pattern of synthesized materials matched the standard spectrum of ZnO (JCPDS card No.

36-1451) with the hexagonal structure. No impurity peak was observed, indicating the single phase of the ZnO

nanoplates [42]. The average crystalline size of ZnO crystals was approximately 10 nm, which was much smaller

than the size of the nanoplates. This result confirmed that the ultrathin porous ZnO nanoplate was polycrystalline.

The composition of the ultrathin porous ZnO nanoplates was studied by EDS (Figure 1[D]), which showed

evidence of Zn and O. The nanoplates were of high quality without any detectable impurities. The content of O

of
and Zn was 53.56 and 46.44 at.%, respectively. This composition was approximate to the stoichiometry of ZnO,

ro
confirming that a high-quality material was obtained [43].

<preferred position of Figure 1>


-p
HRTEM images were used to further study the crystal structure and the crystallinity of the ultrathin

porous ZnO nanoplates (Figure 2). Figure 2(A) shows a TEM image of the ZnO nanoplates, which indicates that
re
the nanoplates were not uniform but highly porous. Some aggregated tiny particles can also be observed in the

HRTEM image, possibly due to fracture of the material during sampling by ultrasonication. Many large pores can
lP

be easily observed in the nanoplates, confirming the porous structure of the material [44]. The porous structure of

the nanoplates was more evident in the higher magnification HRTEM (Figure 2[B]). The inset of Figure 2(B) is a
na

cross-sectional HRTEM image of the synthesized ZnO nanoplates, which reveals that the nanoplates have an

average thickness of 15 nm. The surface of the nanoplates was not as flat as that of the single crystal but rough

due to the formation of nanopores. The HRTEM image of the corresponding area of the nanoplate shown in
ur

Figure 2(C) shows the high crystallinity of the nanoplate where the lattice fringes were visible. However, the

orientation of the lattice fringes was different, suggesting that the nanoplate was composed of different
Jo

nanocrystals. Figure 2(D) displays the enlarged HRTEM image, which shows clear lattice fringes with an

interspacing of 0.52 nm. The distance between lattice fringes is consistent with the gap between (0001) planes of

hexagonal ZnO. This result indicated that the synthesized nanoplate was not a single crystal but polycrystalline

with different orientations of the nanocrystals. However, the preferred growth orientation of the nanocrystals was
[0001] direction [33,45]. This is advantageous because ZnO crystal with exposed facets of {0001} could enhance

the sensor performance [46].

<preferred position of Figure 2>

The porosity and specific surface area of the ultrathin porous ZnO nanoplates were determined by BET

measurement [33]. The isotherm nitrogen adsorption/desorption (Figure 3[A]) displays the characteristics of type

IV porous materials [47]. There are two peaks at approximately 2.5 and 100 nm in the pore-size distribution of the

ZnO nanoplates (inset of Figure 3[A]). The first pore (2.5 nm) was attributed to the nanopores inside the

nanoplates, whereas the second pore (100 nm) was formed by the interspaces between the ZnO nanoplates. The

BET surface area of the ultrathin porous ZnO nanoplates was 74.976 m2/g, which was higher than that of the ZnO

of
nest-like architecture (63.46 m2/g) [46] and precipitated ZnO nanosheets (23.88 m2/g) [33]. The large BET surface

area with many pores of the ZnO nanoplates was advantageous for enhancing gas sensitivity [48].

ro
The PL spectrum of the ultrathin porous ZnO nanoplates (Figure 3(B)) displays a small peak at 325 nm,

assigned to the helium–cadmium (He-Cd) excitation laser. In addition, there are two emission peaks centred at
-p
approximately 389 nm and 650 nm. The near-band-edge emission in the UV region is ascribed to the band-to-band

of ZnO nanocrystals (3.188 eV). The broad emission centred at 650 nm was deconvoluted into three peaks at 526,
re
600, and 695 nm. The presence of visible peaks was ascribed to defects in the ZnO nanoplate structure. The peak
lP

at 526 nm was attributed to OZn, whereas yellow (600 nm), and orange (695 nm) emissions result from oxygen

interstitials [49,50]. Herein, the large amount of oxygen interstitials in ZnO is expected to yield a higher sensor

response [42].
na

In addition to the PL data, the Raman spectrum of the porous ZnO nanoplates was also studied, and the

data is shown in Figure 3(C). Three peaks can be observed at 150, 385, and 1060 cm−1, which correspond to the
ur

optical phonon 2E2(L), A1(TO), and A1(TO)+ E2L modes of hexagonal ZnO, respectively [51]. The 2E2L mode at

approximately 150 cm−1 is the strongest, indicating a high level of defects relating to the Zn sublattice vibration
Jo

[52]. The present of multi-phonon scattering modes at 385 and 1060 cm−1 in our result is similar to that reported

in ref. [53].

<preferred position of Figure 3>


The gas sensor based on the ultrathin porous ZnO nanoplates was tested for detecting acetone, ethanol,

methanol, toluene and NH3 gases at temperatures ranging from 350 to 450 °C. Figure 4(A) displays the transient

response versus time of the sensor to acetone (50–125 ppm) measured at different working temperatures. The base

resistance of the sensor in air decreased from 11 k to about 10 k with increasing temperature from 350 to 450

°C. This was caused by the thermal energy, which activates the electrons from the donor levels to the conduction

band and decreases the based resistance of ZnO. At a given temperature, the sensor resistance rapidly decreased

upon exposure to acetone, and reached the saturation value within a minute, indicating the fast response speed.

The result is consistent with recent reports on the acetone sensing behaviour of ZnO, where the decrease of sensor

resistance upon exposure to acetone was ascribed to the reducing behaviour of acetone and the n-type

of
semiconducting properties of ZnO. After stopping the flow of acetone and refreshing with air, the sensor resistance

recovered to the initial value, confirming the good recovery characteristics of the sensor. Figure 4(B) depicts the

ro
dependence of sensor response (Ra/Rg) as a function of acetone concentration at all measured temperatures. At a

given measured acetone concentration, the sensor response was maximal at a working temperature of 450 °C in

-p
the measured range. A decrease in working temperature leads to a decrease in sensor response. For example, the

response value to 50 ppm acetone is 11.64 at 450 °C but it is only 7.8 at 350 °C. This means that the reaction
re
between acetone molecules and the surface of porous ZnO nanoplates required high activation energy. Our result

is consistent with the report by Wongrat et. al [54], where ZnO nanowires exhibited the maximum response to
lP

acetone at a working temperature of 425 °C. Herein, at the working temperature of 450 °C, the sensor response

increased nearly linearly from 11.64 to 20 with increasing acetone concentration from 50 to 125 ppm. The porous

ZnO nanoplates prepared in our study exhibited a better acetone sensing performance than that of Sn-doped ZnO
na

ultrathin nanosheets [38]. Comparing with recent reports on acetone sensors, the sensitivity of the ultrathin porous

ZnO nanoplates was very high compared to Sn-doped ZnO ultrathin nanosheets, which exhibited a response of
ur

5.55 for 200 ppm acetone at 320 °C [38], porous 1% Au@ZnO nanoplates, which exhibited a response of 32 for

100 ppm acetone at 400 °C [34], coral-like ZnO, which exhibited a response of 12 for 100 ppm acetone at 360 °C
Jo

[45], and the nest-like ZnO architecture, which exhibited a response of 11 for 1000 ppm acetone at 105 °C [46].

The response and recovery times of the sensor were calculated from the transient resistance versus time

when the flow was switched from air to gas and back to air. Figure 4(C) shows the response and recovery time of

the sensor measured at 450 °C. At a low acetone concentration of 50 ppm, the response and recovery time of the
sensor was 23 and 637 s, respectively. However, the response time decreased to 7 s and recovery increased to 550

s when the acetone concentration increased from 50 to 125 ppm. The recovery time was much longer than the

response time, possibly due to the strong interaction between acetone molecules and the surface of ZnO. The long

recovery time was also caused by the large volume of the test chamber, which required a long time to refresh the

chamber. The decrease of response time with increased acetone concentration can be explained as follows. When

increasing the acetone concentration, more acetone molecules can come to adsorb on the surface of ZnO and

rapidly occupy the adsorption sites. Because the number of available adsorption sites on the surface of the sensing

layer is fixed at a given working temperature, increasing the acetone concentration leads to reduced response time

[55]. By contrast, the recovery time of the sensor increased with an increase of acetone concentration because at

the dynamic equilibrium condition, the number of acetone molecules adsorbed on the surface of the sensing layer

of
is high with high concentration; thus more adsorbed molecules require a longer time to desorb. With an increase

ro
of working temperature, the response time and recovery time of the sensor decreased (data not shown). The

recovery time decreased with increased working temperature due to the rapid acetone desorption with the

-p
acceleration of thermal energy. Despite the long recovery time of the sensor, the device still exhibited 100%

recovery characteristics, where the sensor resistance recovered to the base resistance after refreshing the chamber.
re
This result confirmed the physical adsorption of acetone molecules on the surface of ZnO, thus enabling the

reusability of the sensor [46].


lP

<preferred position of Figure 4>


na

We also tested the response of the sensor to low acetone concentrations of subppm levels to apply it in

diabetes diagnosis [56]. Figure 5[A] shows the dynamic sensing response of the ultrathin porous ZnO nanoplates
ur

to 2.510 ppm acetone at 450 °C. The sensor could detect acetone at low concentrations down to 2.5 ppm. The

sensor response increased from 1.8 to 2.0 with an increase of acetone concentration from 2.5 to 10 ppm
Jo

(Figure 5[B]). The response increased as the acetone concentration increased, with a slope of 0.022. We calculated

the lower detection limit of the sensor to be about 45 ppb [56], which is lower than the threshold concentration

(180 ppb) of acetone in breath analysis, indicating the potential application of the sensor for diagnosing diabetes

mellitus. Stability and selectivity are essential aspects of practical applications. Figure 5[C] showed ten pulse

exposures of the sensor to 125 ppm acetone at 450 °C. A negligible response reduction was observed after the test,
indicating the good stability of the sensor. This is the result of the high crystallinity of the ZnO nanoplates. The

selectivity of the ultrathin porous ZnO nanoplate sensor to acetone, ethanol, methanol, NH3 and toluene was tested

at 450 °C. Figure 5(D) shows the response values to different gases, including acetone, ethanol, methanol, NH3

and toluene. The sensor presented the best sensitivity to acetone, followed by ethanol and methanol; the response

values were 20, 9.0 and 4.0, respectively. The sensor presented a negligible response to NH3 and toluene gases.

Those data confirmed that the sensor had good stability and selectivity for detecting acetone at sub-ppm

concentrations.

<preferred position of Figure 5>

Humidity was also present in the exhaled breath; thus, we tested the response of the sensor to 125 ppm

of
acetone in different values of relative humidity (RH), as shown in Figure 6. Figure 6(A) demonstrates that the RH

significantly influenced the base resistance of the sensor and its response. At an operating temperature of 450 °C,

ro
the base resistance was 10, 19 and 36 k for 10, 60 and 80% RH, respectively. The base resistance of the sensor

shifted to higher values with an increase of RH in the measured range, indicating that the adsorbed water molecule
-p
donated free electrons to the conduction band of ZnO to increase the base resistance. However, the presence of

humidity seems not to influence the response time and recovery time of the sensor, where the sensor exhibited
re
100% recovery in different RH values. Figure 6(B) showed the response values of the sensor measured at different

RH. The response value to 125 ppm acetone was 20, 13.6, and 11.86 in 10, 60 and 80% RH, respectively. The
lP

humidity in exhaled breath is relatively high, about 80–90%. The ambient humidity influenced the acetone

response, but at high humidities of 60%–80% RH, the variation in acetone response was ignorable. Therefore, the

sensor is suitable for application in exhaled breath analysis without dehumidification.


na

<preferred position of Figure 6>

We compared the acetone sensing response of the ultrathin porous ZnO sensor with other materials, as
ur

shown in Table 1. The ultrathin porous ZnO nanoplates exhibited better acetone gas-sensing properties than those

of 0.5 wt% Co-doped ZnO nanofibers, porous ZnO/ZnCo2O4 hollow spheres, Ag-ZnO NCs, GO/ZnO-NR/GO,
Jo

ZnO rods, CuO nanosheets, nanoporous TiO2 and Co3O4 crossed nanosheets. However, its performance is less

than that of the WO3-C3N4 nanosheet. The sensing performance was believed to be dependent on the materials,

their morphology and composition. The ultrathin porous ZnO nanoplates have a large specific surface area (74.976

m2/g), ultrathin thickness (15 nm) and large pore size (2.5 nm), thus possessed a high response value. The porous
ZnO nanoplates exhibited a higher sensitivity of 20/125 ppm, compared the pure ZnO rods with 4.5/100 ppm.

However, the ultrathin porous ZnO nanoplates required a high operating temperature of 450 °C, compared with

the 300 °C of the ZnO rods. This means the ultrathin porous ZnO nanoplates consume more energy for device

operation, thus limiting its potential application in portable sensing systems. However, such parameters can be

controlled using surface catalyst modification and/or heterojunction [34].

Table 1: Acetone response of different materials for comparison

Metal oxide sensors C (ppm) Response T (°C) Ref

0.5 wt% Co-doped ZnO nanofibers 100 16 360 [57]

porous ZnO/ZnCo2O4 hollow 100 7.5 275 [8]

of
spheres

ro
GO/ZnO-NR/GO 200 12.52 450 [58]

Ag–ZnO NCs 100 15 250 [16]

ZnO rods 100 4.5


-p 300 [59]
re
CuO nanosheets 200 7.725 370 [60]

TiO2 nanoporous 500 25.97 370 [61]


lP

Co3O4 crossed nanosheet (CNS) 1000 36.5 110 [62]

WO3-C3N4 nanosheet 100 35 340 [63]


na

Porous ZnO nanoplates 125 20 450 This work


ur

Density functional theory calculation [64] pointed out that upon exposure to acetone vapour at

approximately 430 °C, the pre-adsorbed oxygen species (O−, O2−) were dissociated from the surface of ZnO and
Jo

replaced by acetone molecules. In addition, the calculation also found that the oxygen extracted more electrons

from the surface than acetone. Thus, the gas sensing mechanism of the ZnO nanoplates can be explained by the

band bending model, in which interaction between the analytic gas molecules and the surface of the sensitive

material modulates the electron-depleted region of the ZnO crystals and leads to the change in resistance of the
sensor [56]. Figure 7 shows a proposed gas sensing mechanism of the ultrathin porous ZnO nanosheets. The sensor

was tested in the air as a reference; thus, the oxygen in ambient air played an important role in the gas sensing

mechanism because, during measurement in air, oxygen molecules were adsorbed on the surface of the ZnO

nanosheets, according to Eqs. (1)–(4)) [65]:

O2 (gas) → O2 (ads), (1)

O2 (ads) + e− → O2−, (2)

O2 (ads) + 2e− → 2O−, (3)

O2 (ads) + 4e− → 2O2−, (4)

The adsorbed oxygen species extract electrons from the conduction of ZnO to form a thick depletion

region on the surface of the ZnO nanoplate (Figure 7[A]). Due to the thin thickness of the ZnO nanoplates with

of
many pores, the oxygen adsorbed and depleted the total surface of the ZnO nanoplates, and thus, the sensor has

ro
high base resistance in air. Upon introduction of acetone, the acetone molecules reacted with the preabsorbed

oxygen species and released electrons back to the conduction band of ZnO [46], as described by Equations (5–6):

-p
CH3COCH3(gas) + 8O−(ads) → 3CO2 + 3H2O + 8e−

CH3COCH3(gas) + 8O2− (ads) → 3CO2 + 3H2O + 16e− (6)


(5)
re
The released electrons decrease the thickness of the depletion region on the surface of the ZnO nanoplate,

as shown in Figure 7(B). As a result, the exposure to acetone gas decreases the sensor resistance compared to the
lP

base value. Not only is there a significant change in the depleted region to maximize the sensing performance, but

the porous structure of the ZnO nanoplates also enables the effective diffusion of the gas molecules to enhance the

response speed. As a result, the ultrathin porous ZnO nanoplates exhibited remarkably high sensitivity [19]. The
na

high sensitivity of the sensor obtained in this study is comparable with that of graphene–Ag loaded ZnO

nanocomposites, which showed a response of 83 for 1000 ppm acetone at 175 °C [66].
ur

<preferred position of Figure 6>

4. Conclusion
Jo

In conclusion, the ultrathin porous ZnO nanoplates were prepared for acetone gas sensing by an

environmentally friendly hydrothermal method without using any surfactant or organic solvent. The prepared ZnO

materials possess a nanoplate morphology with a porous structure and thin thickness of approximately 15 nm. The

sensor fabricated from the ultrathin porous ZnO nanoplates exhibits high sensitivity, selectivity and stability to
low-acetone concentrations (2.5–125 ppm) at a working temperature of 450 °C. The response to 2.5 ppm acetone

is approximately 1.82, which is superior to other reported acetone sensors. The detection limit of the sensor to

acetone at 450 °C is 45 ppb, which is lower than the threshold limit concentration of acetone in diagnosing diabetes

through breath analysis. This result indicates that the ultrathin porous ZnO nanoplates are potential materials for

achieving excellent acetone-sensing performance with detection at the ppb level, enabling application in breath

analysis.

Novelty statement:

of
High performance acetone gas sensor is potential for application in diabetes diagnose via
exhaled breath analysis. Here, we used a urea mediated hydrothermal method to synthesize

ro
ultrathin porous ZnO nanoplates. The synthesized ZnO nanoplate has a pore size of 2.5 nm and
a thickness of 15 nm, which showed excellent acetone sensing performance with detection limit
of 45 ppb.
-p
re
Declaration of interests
lP

The authors declare that they have no known competing financial interests or
personal relationships that could have appeared to influence the work reported in
this paper.
na

Acknowledgement
ur

This work was funded by the Vietnam National Foundation for Science and Technology Development

(NAFOSTED) under grant number: 14/2020/TN.


Jo
References

[1] M.J. Wilde, R.L. Cordell, D. Salman, B. Zhao, W. Ibrahim, L. Bryant, D. Ruszkiewicz, A. Singapuri, R.C.

Free, E.A. Gaillard, C. Beardsmore, C.L.P. Thomas, C.E. Brightling, S. Siddiqui, P.S. Monks, Breath

analysis by two-dimensional gas chromatography with dual flame ionisation and mass spectrometric

detection – Method optimisation and integration within a large-scale clinical study, J. Chromatogr. A. 1594

(2019) 160–172. doi:10.1016/j.chroma.2019.02.001.

[2] M. Zhou, R. Sharma, H. Zhu, Z. Li, J. Li, S. Wang, E. Bisco, J. Massey, A. Pennington, M. Sjoding, R.P.

Dickson, P. Park, R. Hyzy, L. Napolitano, C.E. Gillies, K.R. Ward, X. Fan, Rapid breath analysis for acute

respiratory distress syndrome diagnostics using a portable two-dimensional gas chromatography device,

of
Anal. Bioanal. Chem. 411 (2019) 6435–6447. doi:10.1007/s00216-019-02024-5.

[3] A. Rydosz, Sensors for Enhanced Detection of Acetone as a Potential Tool for Noninvasive Diabetes

ro
Monitoring, Sensors. 18 (2018) 2298. doi:10.3390/s18072298.

[4] -p
M. Righettoni, A. Amann, S.E. Pratsinis, Breath analysis by nanostructured metal oxides as chemo-

resistive gas sensors, Mater. Today. 18 (2015) 163–171. doi:10.1016/j.mattod.2014.08.017.


re
[5] C. Stönner, A. Edtbauer, J. Williams, Real-world volatile organic compound emission rates from seated

adults and children for use in indoor air studies, Indoor Air. 28 (2018) 164–172. doi:10.1111/ina.12405.
lP

[6] A.H. Jalal, F. Alam, S. Roychoudhury, Y. Umasankar, N. Pala, S. Bhansali, Prospects and Challenges of

Volatile Organic Compound Sensors in Human Healthcare, ACS Sensors. 3 (2018) 1246–1263.
na

doi:10.1021/acssensors.8b00400.

[7] A. De Vincentis, G. Pennazza, M. Santonico, U. Vespasiani-Gentilucci, G. Galati, P. Gallo, C. Vernile, C.


ur

Pedone, R. Antonelli Incalzi, A. Picardi, Breath-print analysis by e-nose for classifying and monitoring

chronic liver disease: a proof-of-concept study, Sci. Rep. 6 (2016) 25337. doi:10.1038/srep25337.
Jo

[8] C. Peng, J. Guo, W. Yang, C. Shi, M. Liu, Y. Zheng, J. Xu, P. Chen, T. Huang, Y. Yang, Synthesis of

three-dimensional flower-like hierarchical ZnO nanostructure and its enhanced acetone gas sensing

properties, J. Alloys Compd. 654 (2016) 371–378. doi:10.1016/j.jallcom.2015.09.120.

[9] C. Su, L. Zhang, Y. Han, X. Chen, S. Wang, M. Zeng, N. Hu, Y. Su, Z. Zhou, H. Wei, Z. Yang, Glucose-
assisted synthesis of hierarchical flower-like Co3O4 nanostructures assembled by porous nanosheets for

enhanced acetone sensing, Sensors Actuators B Chem. 288 (2019) 699–706.

doi:10.1016/j.snb.2019.03.004.

[10] P. Srinivasan, A.J. Kulandaisamy, G.K. Mani, K.J. Babu, K. Tsuchiya, J.B.B. Rayappan, Development of

an acetone sensor using nanostructured Co 3 O 4 thin films for exhaled breath analysis, RSC Adv. 9 (2019)

30226–30239. doi:10.1039/C9RA04230J.

[11] J.-W. Yoon, J.-H. Lee, Toward breath analysis on a chip for disease diagnosis using semiconductor-based

chemiresistors: recent progress and future perspectives, Lab Chip. 17 (2017) 3537–3557.

doi:10.1039/C7LC00810D.

of
[12] U.T. Nakate, R. Ahmad, P. Patil, Y.T. Yu, Y.-B. Hahn, Ultra thin NiO nanosheets for high performance

hydrogen gas sensor device, Appl. Surf. Sci. 506 (2020) 144971. doi:10.1016/j.apsusc.2019.144971.

ro
[13] A. Staerz, U. Weimar, N. Barsan, Understanding the Potential of WO3 Based Sensors for Breath Analysis,

[14]
Sensors. 16 (2016) 1815. doi:10.3390/s16111815. -p
G. Atanasova, A.O. Dikovska, T. Dilova, B. Georgieva, G.V. Avdeev, P. Stefanov, N.N. Nedyalkov,
re
Metal-oxide nanostructures produced by PLD in open air for gas sensor applications, Appl. Surf. Sci. 470

(2019) 861–869. doi:10.1016/j.apsusc.2018.11.178.


lP

[15] R. Li, Y. Zhou, M. Sun, Z. Gong, Y. Guo, X. Yin, F. Wu, W. Ding, Gas sensing selectivity of oxygen-

regulated SnO2 films with different microstructure and texture, J. Mater. Sci. Technol. 35 (2019) 2232–
na

2237. doi:10.1016/j.jmst.2019.06.005.

[16] J. Zhang, H. Lu, L. Zhang, D. Leng, Y. Zhang, W. Wang, Y. Gao, H. Lu, J. Gao, G. Zhu, Z. Yang, C.
ur

Wang, Metal–organic framework-derived ZnO hollow nanocages functionalized with nanoscale Ag

catalysts for enhanced ethanol sensing properties, Sensors Actuators B Chem. 291 (2019) 458–469.
Jo

doi:10.1016/j.snb.2019.04.058.

[17] A. Mirzaei, J.-H. Kim, H.W. Kim, S.S. Kim, Resistive-based gas sensors for detection of benzene, toluene

and xylene (BTX) gases: a review, J. Mater. Chem. C. 6 (2018) 4342–4370. doi:10.1039/C8TC00245B.

[18] R. Gao, X. Cheng, S. Gao, X. Zhang, Y. Xu, H. Zhao, L. Huo, Highly selective detection of saturated
vapors of abused drugs by ZnO nanorod bundles gas sensor, Appl. Surf. Sci. 485 (2019) 266–273.

doi:10.1016/j.apsusc.2019.04.189.

[19] L. Van Duy, N.H. Hanh, D.N. Son, P.T. Hung, C.M. Hung, N. Van Duy, N.D. Hoa, N. Van Hieu, Facile

Hydrothermal Synthesis of Two-Dimensional Porous ZnO Nanosheets for Highly Sensitive Ethanol

Sensor, J. Nanomater. 2019 (2019) 1–7. doi:10.1155/2019/4867909.

[20] Y. Park, R. Yoo, S. ryull Park, J.H. Lee, H. Jung, H.-S. Lee, W. Lee, Highly sensitive and selective

isoprene sensing performance of ZnO quantum dots for a breath analyzer, Sensors Actuators B Chem. 290

(2019) 258–266. doi:10.1016/j.snb.2019.03.118.

[21] L. Zhu, W. Zeng, Y. Li, New insight into gas sensing property of ZnO nanorods and nanosheets, Mater.

of
Lett. 228 (2018) 331–333. doi:10.1016/j.matlet.2018.06.049.

ro
[22] M. Yang, S. Zhang, F. Qu, S. Gong, C. Wang, L. Qiu, M. Yang, W. Cheng, High performance acetone

sensor based on ZnO nanorods modified by Au nanoparticles, J. Alloys Compd. 797 (2019) 246–252.

[23]
doi:10.1016/j.jallcom.2019.05.101. -p
S. Agarwal, P. Rai, E.N. Gatell, E. Llobet, F. Güell, M. Kumar, K. Awasthi, Gas sensing properties of
re
ZnO nanostructures (flowers/rods) synthesized by hydrothermal method, Sensors Actuators B Chem. 292

(2019) 24–31. doi:10.1016/j.snb.2019.04.083.


lP

[24] L. Song, H. Yue, H. Li, L. Liu, Y. Li, L. Du, H. Duan, N.I. Klyui, Hierarchical porous ZnO microflowers

with ultra-high ethanol gas-sensing at low concentration, Chem. Phys. Lett. 699 (2018) 1–7.
na

doi:10.1016/j.cplett.2018.03.021.

[25] N.S. Ramgir, M. Kaur, P.K. Sharma, N. Datta, S. Kailasaganapathi, S. Bhattacharya, A.K. Debnath, D.K.
ur

Aswal, S.K. Gupta, Ethanol sensing properties of pure and Au modified ZnO nanowires, Sensors Actuators

B Chem. 187 (2013) 313–318. doi:10.1016/j.snb.2012.11.079.


Jo

[26] J.-H. Kim, A. Mirzaei, H. Woo Kim, S.S. Kim, Combination of Pd loading and electron beam irradiation

for superior hydrogen sensing of electrospun ZnO nanofibers, Sensors Actuators B Chem. 284 (2019) 628–

637. doi:10.1016/j.snb.2018.12.120.

[27] J. Cui, L. Shi, T. Xie, D. Wang, Y. Lin, UV-light illumination room temperature HCHO gas-sensing
mechanism of ZnO with different nanostructures, Sensors Actuators B Chem. 227 (2016) 220–226.

doi:10.1016/j.snb.2015.12.010.

[28] Y. Wang, M. Yang, W. Liu, L. Dong, D. Chen, C. Peng, Gas sensors based on assembled porous graphene

multilayer frameworks for DMMP detection, J. Mater. Chem. C. 7 (2019) 9248–9256.

doi:10.1039/C9TC02299F.

[29] Z. Dai, T. Liang, J.-H. Lee, Gas sensors using ordered macroporous oxide nanostructures, Nanoscale Adv.

1 (2019) 1626–1639. doi:10.1039/C8NA00303C.

[30] M. Chen, Z. Wang, D. Han, F. Gu, G. Guo, Porous ZnO Polygonal Nanoflakes: Synthesis, Use in High-

Sensitivity NO 2 Gas Sensor, and Proposed Mechanism of Gas Sensing, J. Phys. Chem. C. 115 (2011)

of
12763–12773. doi:10.1021/jp201816d.

ro
[31] S. Leonardi, Two-Dimensional Zinc Oxide Nanostructures for Gas Sensor Applications, Chemosensors. 5

(2017) 17. doi:10.3390/chemosensors5020017.

[32]
-p
C. Gu, H. Huang, J. Huang, Z. Jin, H. Zheng, N. Liu, M. Li, J. Liu, F. Meng, Chlorobenzene sensor based

on Pt-decorated porous single-crystalline ZnO nanosheets, Sensors Actuators A Phys. 252 (2016) 96–103.
re
doi:10.1016/j.sna.2016.11.004.

[33] S.-M. Li, L.-X. Zhang, M.-Y. Zhu, G.-J. Ji, L.-X. Zhao, J. Yin, L.-J. Bie, Acetone sensing of ZnO
lP

nanosheets synthesized using room-temperature precipitation, Sensors Actuators B Chem. 249 (2017)

611–623. doi:10.1016/j.snb.2017.04.007.
na

[34] Z. Feng, Y. Ma, V. Natarajan, Q. Zhao, X. Ma, J. Zhan, In-situ generation of highly dispersed Au

nanoparticles on porous ZnO nanoplates via ion exchange from hydrozincite for VOCs gas sensing,
ur

Sensors Actuators B Chem. 255 (2018) 884–890. doi:10.1016/j.snb.2017.08.138.

[35] F. Meng, J. Yin, Y.-Q. Duan, Z.-H. Yuan, L.-J. Bie, Co-precipitation synthesis and gas-sensing properties
Jo

of ZnO hollow sphere with porous shell, Sensors Actuators B Chem. 156 (2011) 703–708.

doi:10.1016/j.snb.2011.02.022.

[36] Y. Zhu, Y. Wang, G. Duan, H. Zhang, Y. Li, G. Liu, L. Xu, W. Cai, In situ growth of porous ZnO

nanosheet-built network film as high-performance gas sensor, Sensors Actuators B Chem. 221 (2015)
350–356. doi:10.1016/j.snb.2015.06.115.

[37] N.T. Phuong Nhung, P. Van Tong, C.M. Hung, N. Van Duy, N.V. Chien, N. Van Vinh, N.T. Tuyen, N.D.

Hoa, Nanoporous ZnO nanostructure synthesis by a facile method for superior sensitivity ethanol sensor

applications, RSC Adv. 6 (2016) 64215–64218. doi:10.1039/C6RA11531D.

[38] Y. Al-Hadeethi, A. Umar, S.H. Al-Heniti, R. Kumar, S.H. Kim, X. Zhang, B.M. Raffah, 2D Sn-doped

ZnO ultrathin nanosheet networks for enhanced acetone gas sensing application, Ceram. Int. 43 (2017)

2418–2423. doi:10.1016/j.ceramint.2016.11.031.

[39] N.D. Hoa, C.M. Hung, N. Van Duy, N. Van Hieu, Nanoporous and crystal evolution in nickel oxide

nanosheets for enhanced gas-sensing performance, Sensors Actuators B Chem. 273 (2018) 784–793.

of
doi:10.1016/j.snb.2018.06.095.

ro
[40] N. Van Hieu, N. Van Duy, P.T. Huy, N.D. Chien, Inclusion of SWCNTs in Nb/Pt co-doped TiO2 thin-

film sensor for ethanol vapor detection, Phys. E Low-Dimensional Syst. Nanostructures. 40 (2008) 2950–

[41]
2958. doi:10.1016/j.physe.2008.02.018. -p
R. Bo, N. Nasiri, H. Chen, D. Caputo, L. Fu, A. Tricoli, Low-Voltage High-Performance UV
re
Photodetectors: An Interplay between Grain Boundaries and Debye Length, ACS Appl. Mater. Interfaces.

9 (2017) 2606–2615. doi:10.1021/acsami.6b12321.


lP

[42] C.T. Quy, N.X. Thai, N.D. Hoa, D.T. Thanh Le, C.M. Hung, N. Van Duy, N. Van Hieu, C 2 H 5 OH and

NO 2 sensing properties of ZnO nanostructures: correlation between crystal size, defect level and sensing
na

performance, RSC Adv. 8 (2018) 5629–5639. doi:10.1039/C7RA13702H.

[43] V.T. Duoc, D.T.T. Le, N.D. Hoa, N. Van Duy, C.M. Hung, H. Nguyen, N. Van Hieu, New Design of ZnO
ur

Nanorod- and Nanowire-Based NO 2 Room-Temperature Sensors Prepared by Hydrothermal Method, J.

Nanomater. 2019 (2019) 1–9. doi:10.1155/2019/6821937.


Jo

[44] N.D. Hoa, N. Van Duy, S.A. El-Safty, N. Van Hieu, Meso-/Nanoporous Semiconducting Metal Oxides

for Gas Sensor Applications, J. Nanomater. 2015 (2015) 1–14. doi:10.1155/2015/972025.

[45] L. Zhu, W. Zeng, H. Ye, Y. Li, Volatile organic compound sensing based on coral rock-like ZnO, Mater.

Res. Bull. 100 (2018) 259–264. doi:10.1016/j.materresbull.2017.12.043.


[46] C. Li, H. Zhou, S. Yang, L. Wei, Z. Han, Y. Zhang, H. Pan, Preadsorption of O 2 on the Exposed (001)

Facets of ZnO Nanostructures for Enhanced Sensing of Gaseous Acetone, ACS Appl. Nano Mater. 2

(2019) 6144–6151. doi:10.1021/acsanm.9b00942.

[47] F. Meng, N. Hou, S. Ge, B. Sun, Z. Jin, W. Shen, L. Kong, Z. Guo, Y. Sun, H. Wu, C. Wang, M. Li,

Flower-like hierarchical structures consisting of porous single-crystalline ZnO nanosheets and their gas

sensing properties to volatile organic compounds (VOCs), J. Alloys Compd. 626 (2015) 124–130.

doi:10.1016/j.jallcom.2014.11.175.

[48] C.M. Hung, N.D. Hoa, N. Van Duy, N. Van Toan, D.T.T. Le, N. Van Hieu, Synthesis and gas-sensing

characteristics of α-Fe 2 O 3 hollow balls, J. Sci. Adv. Mater. Devices. 1 (2016) 45–50.

of
doi:10.1016/j.jsamd.2016.03.003.

[49] Y.-H. Liu, S.-J. Young, L.-W. Ji, S.-J. Chang, Ga-Doped ZnO Nanosheet Structure-Based Ultraviolet

ro
Photodetector by Low-Temperature Aqueous Solution Method, IEEE Trans. Electron Devices. 62 (2015)

2924–2927. doi:10.1109/TED.2015.2457441.

[50]
-p
Y. Zhu, H. Yang, F. Sun, X. Wang, Controllable Growth of Ultrathin P-doped ZnO Nanosheets, Nanoscale
re
Res. Lett. 11 (2016) 175. doi:10.1186/s11671-016-1379-8.

[51] I. Musa, N. Qamhieh, S.T. Mahmoud, Synthesis and length dependent photoluminescence property of zinc
lP

oxide nanorods, Results Phys. 7 (2017) 3552–3556. doi:10.1016/j.rinp.2017.09.035.

[52] H. Fukushima, T. Kozu, H. Shima, H. Funakubo, H. Uchida, T. Katoda, K. Nishida, Evaluation of oxygen
na

vacancy in ZnO using Raman spectroscopy, in: 2015 Jt. IEEE Int. Symp. Appl. Ferroelectr. (ISAF), Int.

Symp. Integr. Funct. (ISIF), Piezoelectric Force Microsc. Work., IEEE, 2015: pp. 28–31.

doi:10.1109/ISAF.2015.7172660.
ur

[53] M.T. Dieng, B.D. Ngom, P.D. Tall, M. Maaza, Biosynthesis of Zn5(CO3)2(OH)6 from Arachis Hypogaea
Jo

Shell (Peanut Shell) and Its Conversion to ZnO Nanoparticles, Am. J. Nanomater. Vol. 7, 2019, Pages 1-

9. 7 (2019) 1–9. doi:10.12691/AJN-7-1-1.

[54] E. Wongrat, N. Chanlek, C. Chueaiarrom, W. Thupthimchun, B. Samransuksamer, S. Choopun, Acetone

gas sensors based on ZnO nanostructures decorated with Pt and Nb, Ceram. Int. 43 (2017) S557–S566.

doi:10.1016/j.ceramint.2017.05.296.
[55] H. Chen, J. Hu, G.-D. Li, Q. Gao, C. Wei, X. Zou, Porous Ga–In Bimetallic Oxide Nanofibers with

Controllable Structures for Ultrasensitive and Selective Detection of Formaldehyde, ACS Appl. Mater.

Interfaces. 9 (2017) 4692–4700. doi:10.1021/acsami.6b13520.

[56] N.H. Hanh, L. Van Duy, C.M. Hung, N. Van Duy, Y.-W. Heo, N. Van Hieu, N.D. Hoa, VOC gas Sensor

based on Hollow Cubic Assembled Nanocrystal Zn2SnO4 for Breath Analysis, Sensors Actuators A Phys.

(2020) 111834. doi:10.1016/j.sna.2020.111834.

[57] L. Liu, S. Li, J. Zhuang, L. Wang, J. Zhang, H. Li, Z. Liu, Y. Han, X. Jiang, P. Zhang, Improved selective

acetone sensing properties of Co-doped ZnO nanofibers by electrospinning, Sensors Actuators B Chem.

155 (2011) 782–788. doi:10.1016/j.snb.2011.01.047.

of
[58] B.A. Vessalli, C.A. Zito, T.M. Perfecto, D.P. Volanti, T. Mazon, ZnO nanorods/graphene oxide sheets

prepared by chemical bath deposition for volatile organic compounds detection, J. Alloys Compd. 696

ro
(2017) 996–1003. doi:10.1016/j.jallcom.2016.12.075.

[59] -p
X. Yang, S. Zhang, Q. Yu, L. Zhao, P. Sun, T. Wang, F. Liu, X. Yan, Y. Gao, X. Liang, S. Zhang, G. Lu,

One step synthesis of branched SnO2/ZnO heterostructures and their enhanced gas-sensing properties,
re
Sensors Actuators B Chem. 281 (2019) 415–423. doi:10.1016/j.snb.2018.10.138.

[60] A. Umar, A.A. Alshahrani, H. Algarni, R. Kumar, CuO nanosheets as potential scaffolds for gas sensing
lP

applications, Sensors Actuators, B Chem. 250 (2017) 24–31. doi:10.1016/j.snb.2017.04.062.

[61] N. Chen, Y. Li, D. Deng, X. Liu, X. Xing, X. Xiao, Y. Wang, Acetone sensing performances based on
na

nanoporous TiO2 synthesized by a facile hydrothermal method, Sensors Actuators B Chem. 238 (2017)

491–500. doi:10.1016/j.snb.2016.07.094.
ur

[62] Z. Zhang, L. Zhu, Z. Wen, Z. Ye, Controllable synthesis of Co3O4 crossed nanosheet arrays toward an

acetone gas sensor, Sensors Actuators B Chem. 238 (2017) 1052–1059. doi:10.1016/j.snb.2016.07.154.
Jo

[63] D. Wang, S. Huang, H. Li, A. Chen, P. Wang, J. Yang, X. Wang, J. Yang, Ultrathin WO3 nanosheets

modified by g-C3N4 for highly efficient acetone vapor detection, Sensors Actuators B Chem. 282 (2019)

961–971. doi:10.1016/j.snb.2018.11.138.

[64] M. Sadeghian Lemraski, E. Nadimi, Acetone gas sensing mechanism on zinc oxide surfaces: A first
principles calculation, Surf. Sci. 657 (2017) 96–103. doi:10.1016/j.susc.2016.11.013.

[65] G. Li, Y. Shen, P. Zhou, F. Hao, P. Fang, D. Wei, D. Meng, X. San, Design and application of highly

responsive and selective rGO-SnO2 nanocomposites for NO2 monitoring, Mater. Charact. 163 (2020)

110284. doi:10.1016/j.matchar.2020.110284.

[66] D.Y. Nadargi, R.B. Dateer, M.S. Tamboli, I.S. Mulla, S.S. Suryavanshi, A greener approach towards the

development of graphene–Ag loaded ZnO nanocomposites for acetone sensing applications, RSC Adv. 9

(2019) 33602–33606. doi:10.1039/C9RA06482F.

of
ro
-p
re
lP
na
ur
Jo
Figure caption:

Figure 1. SEM images (A–B), XRD patterns (C), and EDS analysis (D) of the ultrathin porous ZnO nanoplates.

of
ro
-p
re
lP

Figure 2. (A, B) Low- and (C, D) high-magnification TEM images of the ultrathin porous ZnO nanoplates. Inset
na

of (B) shows a cross-sectional TEM image of the synthesized ZnO nanoplates.


ur
Jo
of
ro
-p
re
Figure 3. (A) Nitrogen adsorption-desorption isothermal BET surface area, (B) PL and (C) Raman spectroscopy
lP

of the ultrathin porous ZnO nanoplates


na
ur
Jo
of
ro
-p
re
Figure 4. Acetone-sensing characteristics of the ultrathin porous ZnO nanoplates measured at 350, 400, and
lP

450 °C: (A) transient resistance vs. time, (B) the dependence of response on acetone concentrations, (C) response

and recovery time of the sensor.


na
ur
Jo
of
ro
-p
re
Figure 5. (A) ZnO nanoplate response to low-acetone concentrations at 450 °C, (B) response vs. acetone
lP

concentration, (C) stability of the ZnO sensor to acetone, (D) selectivity of the sensor.
na
ur
Jo
of
ro
-p
re
Figure 6. Dynamic resistance curves (A), the response (B) of the sensor to 125 ppm acetone with different relative
lP

humidity values measured at 450 °C.


na
ur
Jo
of
ro
-p
re
lP

Figure 7. Proposed gas-sensing mechanism: (A) ultrathin porous ZnO nanoplates in air; (B) ultrathin porous ZnO

nanoplates in acetone
na
ur
Jo
Jo
ur
na
lP
re
-p
ro
of

You might also like