You are on page 1of 9

Journal of

Materials Chemistry A
View Article Online
PAPER View Journal | View Issue
Published on 16 November 2020. Downloaded by University of California - Santa Barbara on 5/16/2021 1:30:12 PM.

Individual gas sensor detecting dual exhaled


Cite this: J. Mater. Chem. A, 2020, 8,
biomarkers via a temperature modulated n/p
26004 semiconducting transition†
Xiaxia Xing,a Lingling Du,a Dongliang Feng,a Chen Wang,a Mingshui Yao,b
Xiaohu Huang,c Shixi Zhangd and Dachi Yang *a

The newly-emerged breath sensing detection of disease biomarkers (e.g., H2S and acetone) offers rapid and
noninvasive early diagnosis of diseases (e.g., halitosis and diabetes). Utilizing a single sensor to monitor dual
biomarkers may contribute to both miniaturized size and multi-detection if implemented, but remains
challenging. Here, interconnected BiFeO3/Bi25FeO40 nanoparticles (NPs, 22.5 nm in diameter)
synthesized via combined microwave hydrothermal and annealing methods have been developed to
selectively detect dual biomarkers of halitosis and diabetes. The selective detection has been modulated
by an n/p semiconducting transformation, and has been experimentally observed and theoretically
interpreted, in which electron–hole pairs are modulated due to the synergistic effect of temperature-
dependent adsorbed oxygen molecules and semiconducting band bending. Remarkably, the sensor
prototypes enable the selective detection of both H2S (n-type mode) and acetone (p-type mode)
biomarkers with superior stability and a ppb-level detection limit. Furthermore, practical human breath
Received 22nd September 2020
Accepted 13th November 2020
has been experimentally simulated. Critically, the sensors with a waterproof membrane have been tested
by immersing them into water. Our strategy of the detection of dual exhaled biomarkers by a single gas
DOI: 10.1039/d0ta09321a
sensor may contribute to the integration and miniaturization of sensors, for upcoming intelligent medical
rsc.li/materials-a treatment.

Additionally, to achieve multi-detection and miniaturization,


1. Introduction the challenging issue is utilizing a single gas sensor to selec-
Exhaled biomarker analysis provides a feasible route towards tively monitor two or more exhaled biomarkers.
early diagnosis of serious diseases such as lung cancer, gastric To address such issues of sensing performance, the inte-
cancer and diabetes,1–4 and is thus of great potential in the grated sensing materials, such as semiconductor metal oxides
upcoming intelligent medical treatment eld.5 Traditionally, (SMOs), play crucial roles and are hence widely studied. Basi-
the analysis of biomarkers requires complicated procedures, cally, the SMO sensing mechanism is determined by the resis-
time consumption, expensive facilities and procient exper- tance, in which the target gases are adsorbed onto and nally
tise.6–8 Currently, gas-sensing detection of human breath offers vary the carrier concentration of the SMO. To improve the
rapid, inexpensive and non-invasive early-diagnosis.4,9 Gener- selectivity, the approaches of doping precious metals,10,11 con-
ally, such gas sensors in the detection of biomarkers are structing a catalytically active overlayer12–14 and exploiting the
simultaneously endowed with excellent selectivity, stability and interaction between the target gas and sensing-material
ppb-level detection. Although progress has been made, devel- surface15,16 have been explored. For improving the stability,
oping such advanced sensors remains challenging. the methods of gas ltration,17 assembling moisture-resistant
materials18,19 and applying core/shell nanostructures20,21 have
a
been reported. To enable ppb-level detection, decorating
Tianjin Key Laboratory of Optoelectronic Sensor and Sensing Network Technology,
Department of Electronics, College of Electronic Information and Optical
precious metals,22 constructing heterojunctions23 and light
Engineering, Nankai University, Tianjin 300350, China. E-mail: yangdachi@nankai. assistance24–26 have been frequently utilized.
edu.cn Recently, the behavior of an n/p semiconducting transition
b
Institute for Integrated Cell-Material Sciences, Kyoto University Institute for Advanced in SMO sensing materials has been discovered under gas-
Study, Kyoto University, Yoshida Ushinomiya-cho, Sakyo-ku, Kyoto 606-8501, Japan sensing environments, which can notably be modulated by
c
Institute of Materials Research and Engineering, Agency for Science, Technology and
factors such as temperature,27,28 doping concentration,29 and
Research, 138634, Singapore
d
target gas concentration.30 Essentially, such transition behav-
Shangqiu Municipal Hospital, Shangqiu, 476100, China
† Electronic supplementary information (ESI) available. See DOI:
iors under interaction with the target gas are determined by
10.1039/d0ta09321a semiconducting materials, including pure SMOs,31–33 metallic-

26004 | J. Mater. Chem. A, 2020, 8, 26004–26012 This journal is © The Royal Society of Chemistry 2020
View Article Online

Paper Journal of Materials Chemistry A


Published on 16 November 2020. Downloaded by University of California - Santa Barbara on 5/16/2021 1:30:12 PM.

Fig. 1 The morphological and crystalline characterization of BiFeO3/Bi25FeO40 NPs. (a) SEM image. (b and f) TEM images. (c) The size distribution.
(d) The lattice fringes and (e) the SAED pattern taken from the dashed region in Fig. S4b,† respectively. The separated EDS mappings correspond
to (f1) Bi, (f2) Fe and (f3) O elements, respectively. The scale bar in (e) is 2 nm1.

particle doped SMOs29 or SMO heterojunctions.28,30 Interest- has been studied. Remarkably, the sensors present superior
ingly, bismuth ferrite, as a ferroelectric material, enables both stability and a low detection limit of 100 ppb H2S (Ra/Rg  1 ¼
perovskite (BiFeO3) and sillenite (Bi25FeO40) to be presented,34 21.5%) and 500 ppb acetone (Rg/Ra  1 ¼ 10.0%). Also, the
with which gas sensors are built, showing a specic sensing simulated breath sensing of both H2S and acetone biomarkers
response to acetone,35 SO2 (ref. 36) and NH3.37 Apparently, the is of potential in early-diagnosis of halitosis and diabetes.
internet of things requires that fewer sensors are integrated for Moreover, the sensors with a waterproof membrane were tested
minimized size, however, to endow more functions and by critically immersing them in water for 120 min. Theoreti-
stronger performance.38,39 Being inspired by this, if n/p semi- cally, the n/p semiconducting transition is ascribed to separated
conducting transition of sensing materials, e.g., perovskite- electron–hole pairs of interconnected BiFeO3/Bi25FeO40 NPs,
BiFeO3 and sillenite-Bi25FeO40 hetero-nanostructures, could be regulated by ionization of adsorbed oxygen molecules. As
modulated by such parameters as temperature and geometrical a result, such temperature-modulated n/p transition sensing of
size, individual gas sensor prototypes might be able to selec- BiFeO3/Bi25FeO40 NPs enables the selective detection of H2S (n-
tively detect binary target gases (e.g., H2S and acetone type mode) and acetone (p-type mode) biomarkers, which
biomarkers). contribute to the integration and miniaturization of sensor
In this study, selective detection of binary H2S and acetone prototypes.
biomarkers has been developed via temperature-modulated n/p
transformation of interconnected BiFeO3/Bi25FeO40 nano-
particles (NPs). The BiFeO3/Bi25FeO40 NPs of 22.5 nm diam- 2. Results and discussion
eter were obtained by microwave hydrothermal and subsequent 2.1 Synthesis and characterization of BiFeO3/Bi25FeO40 NPs
annealing methods, as is schematically shown in Fig. S1.†
The BiFeO3/Bi25FeO40 NPs in our case were synthesized by
Detection of H2S (n-type mode) and acetone (p-type mode)
combined microwave hydrothermal treatment of Bi(NO3)3 and
biomarkers via a single BiFeO3/Bi25FeO40 NP sensor prototype
Fe(NO3)3 in alkaline media and subsequent annealing. The

This journal is © The Royal Society of Chemistry 2020 J. Mater. Chem. A, 2020, 8, 26004–26012 | 26005
View Article Online

Journal of Materials Chemistry A Paper


Published on 16 November 2020. Downloaded by University of California - Santa Barbara on 5/16/2021 1:30:12 PM.

Fig. 2 XRD, Mott–Schottky plot and XPS characterization of BiFeO3/Bi25FeO40 NPs. (a) XRD patterns, (b) Mott–Schottky plot and (c) the XPS
survey spectrum of the interconnected BiFeO3/Bi25FeO40 NPs. High-resolution XPS spectra are related to (d) Bi 4f, (e) Fe 2p and (f) O 1s,
respectively.

synthetic details are demonstrated in the Experimental interconnected BiFeO3/Bi25FeO40 NPs, HRTEM and selected
section.† The chemical reactions are summarized as follows:40 area electron diffraction (SAED) analysis were performed on the
dashed region in Fig. S4b.† From Fig. 1d, in the HRTEM image,
Bi(NO)3 + Fe(NO)3 + 6KOH / Bi(OH)3 + Fe(OH)3 + the interplanar spacings of d ¼ 0.184 nm and 0.317 nm agree
6KNO3 (1) with the (521) and (310) planes of Bi25FeO40, respectively.
Meanwhile, that of d ¼ 0.279 nm is assigned to the (110) plane
Bi(OH)3 + Fe(OH)3 / BiFeO3 + 3H2O (2) of BiFeO3. Diffractive spots and rings coexist in the SAED
pattern in Fig. 1e and S4c† (the raw data), which reveals the
2BiFeO3 + 48Bi(OH)3 + O2 / 2Bi25FeO40 + 72H2O (3)
polycrystalline nature of the BiFeO3/Bi25FeO40 NPs, agreeing
In our experiments, the synthetic parameters of microwave with the lattice fringes (Fig. 1d) and XRD (Fig. 2a) analysis. The
hydrothermal treatment, such as temperature, duration and diffractive spots are attributed to excellent crystallization in the
chemical composition of the precursors, are essential to form characterized region. In addition, to identify the chemical
the BiFeO3/Bi25FeO40 NPs. In particular, the annealing condi- components of the interconnected BiFeO3/Bi25FeO40 NPs, the
tions, such as temperature and heating rate, play a critical role. elemental mappings of Fig. 1f are exhibited in Fig. 1f1–f3,
We tried the experiments without annealing, and found the revealing that Bi, Fe and O are distributed uniformly.
presence of Fe2O3 phase (Fig. S2†), which is ascribed to the The XRD pattern in Fig. 2a reveals the phase of inter-
thermal decomposition of Fe(OH)3. Meanwhile, in the process connected BiFeO3/Bi25FeO40 NPs. Accordingly, the diffractive
of calcination at high temperature, Bi(OH)3 rstly decomposes peaks agree well with the two-phase coexistence of BiFeO3
into Bi2O3 and then reacts with Fe2O3 to form BiFeO3, and the (rhombohedral crystal, JPCDS no. 86-1518) and Bi25FeO40 (cubic
mixed phases of BiFeO3 and Bi25FeO40 are nally obtained. crystal, JPCDS no. 46-0416). To investigate the Schottky contacts
Otherwise, if other concentrations of polyethylene glycol (PEG) of the interconnected BiFeO3/Bi25FeO40 NPs, we carried out
and KNO3 were employed, NPs with large diameter and uneven Mott–Schottky analysis by means of a mono-frequency capaci-
size distribution were obtained, as shown in Fig. S3a–c.† tance–voltage (C–V) sweep as shown in Fig. 2b, from which the
For morphological observation of BiFeO3/Bi25FeO40 NPs, plots with a positive slope reveal the n-type semiconducting
SEM and TEM characterizations were conducted as shown in feature.4,41 The surface chemical composition and valence state
Fig. 1a and b. One can see that the NPs are uniform and gran- of the interconnected BiFeO3/Bi25FeO40 NPs are essential for the
ular in appearance, and the average diameter of the NPs has sensing performance, and XPS analysis was thus carried out.
been analysed to be 22.5 nm (Fig. 1c). From closer observation Apparently, the survey spectrum in Fig. 2c presents Bi, Fe and O
in Fig. S4a,† the BiFeO3/Bi25FeO40 NPs overlap and form inter- elements, respectively. Moreover, the Bi 4f spectrum in Fig. 2d
connected NPs. To investigate the crystallization of the shows both peaks located at 158.83 and 164.08 eV, which

26006 | J. Mater. Chem. A, 2020, 8, 26004–26012 This journal is © The Royal Society of Chemistry 2020
View Article Online

Paper Journal of Materials Chemistry A

correspond to the Bi3+ and Bi5+, respectively.42 Similarly, it has the n/p sensing transition may vary with the applied target
been conrmed that the formula of Bi25FeO40 is Bi243+(Bi5+Fe3+) gases, and it exists within a temperature range. Additionally, n-
O40, in which Bi3+ ions occupy the octahedral positions, whereas and p-type mixed modes responses appear, which are marked
the Bi5+ and Fe3+ ions share the tetrahedral positions.42,43 In the with red dashed circles in Fig. 3a–e. As a result, the above
Fe 2p highly-resolved spectrum (Fig. 2e), the binding energies of responses are summarized in Fig. 3f, from which the sensing
responses show the best selectivity to H2S at 200  C and present
Published on 16 November 2020. Downloaded by University of California - Santa Barbara on 5/16/2021 1:30:12 PM.

710.59 eV (Fe 2p3/2) and 724.04 eV (Fe 2p1/2) with a satellite peak
at 719.84 eV agree well with Fe3+.42,44 Meanwhile, Fig. 2f shows n-type mode dominations. Meanwhile, under the p-type sensing
the highly-resolved spectrum of O 1s, from which two tted mode, the best selectivity to acetone is seen at 300  C.
peaks corresponding to lattice oxygen and chemisorbed oxygen To investigate the concentration-dependent n/p transition
are observed at 529.93 and 531.14 eV, respectively. From the sensing, we have tested the sensing responses to 0.1–1 ppm H2S
integrated area, the chemisorbed oxygen is calculated to be and 0.5–2 ppm acetone at the working temperatures of 160–
41.54%. Accordingly, it has been proven that multivalent tran- 400  C (Fig. S5 and S6†), and one can see that the n/p transition
sition metal oxides are able to adsorb more surface oxygen,32 does not occur in this concentration region. Notably, the
suggesting high oxygen adsorption on the surface of the inter- BiFeO3/Bi25FeO40 NP sensor prototypes show n-type sensing of
connected BiFeO3/Bi25FeO40 NPs. 0.1–10 ppm H2S at 200  C (Fig. 4b and f) and p-type sensing of
0.5–10 ppm acetone (Fig. 5b) at 300  C, respectively, and do not
display a concentration-dependent n/p transition below
2.2 Temperature-modulated n/p semiconducting transition
10 ppm. Therefore, the working temperature is the key factor for
As the temperature-dependent n/p transition plays a pivotal role the n/p semiconducting transition of the BiFeO3/Bi25FeO40 NPs
in modulating the sensing behavior of interconnected BiFeO3/ in our cases. In particular, compared to the other n/p sensing
Bi25FeO40 NPs, we elaborately evaluated the real-time resistance transition,27–33 our sensors are able to selectively detect H2S (n-
under operating temperatures from 160  C to 400  C toward mode) and acetone (p-mode) via modulating the n/p transition.
100 ppm formaldehyde, ethanol, toluene and acetone, and
10 ppm H2S (Fig. 3a–e). One can see that the real-time resistance
curves of the BiFeO3/Bi25FeO40 NPs sensor display n-type sem- 2.3 Selective detection of H2S and acetone biomarkers
iconducting features below 300  C toward the above- To gain insight into the sensing performance, the performance
mentioned target gases. The behavior of the n/p semi- of the BiFeO3/Bi25FeO40 NPs sensor toward the H2S biomarker
conducting transition occurs at 260–300  C, except for in an under n-type sensing mode was evaluated rst. From Fig. 4a, we
ethanol atmosphere (300–340  C). In contrast, the sensing observed that the sensor shows a maximum response (Ra/Rg  1
response dominated by a p-type mode was observed from 300  C ¼ 1.95) at the optimal operating temperature of 200  C.
to 400  C toward formaldehyde, toluene, acetone and H2S, and Meanwhile, the real-time response plots of H2S (100 ppb to 1
340–400  C to ethanol. Accordingly, the critical temperature of ppm) in Fig. 4b show an excellent linear relationship between

Fig. 3 Sensing-resistance curves of the BiFeO3/Bi25FeO40 NPs sensor prototype towards various target gases. 100 ppm (a) formaldehyde, (b)
ethanol, (c) toluene and (d) acetone, and (e) 10 ppm H2S at various working temperatures. (f) The summarized response histogram. The relative
humidity (RH) of the above test is 30%.

This journal is © The Royal Society of Chemistry 2020 J. Mater. Chem. A, 2020, 8, 26004–26012 | 26007
View Article Online

Journal of Materials Chemistry A Paper


Published on 16 November 2020. Downloaded by University of California - Santa Barbara on 5/16/2021 1:30:12 PM.

Fig. 4 Under n-type mode, the sensing performance of the BiFeO3/Bi25FeO40 NPs sensor prototypes toward H2S at 200  C and RH of 30%. (a)
The responses to 1 ppm at various temperatures. (b) The real-time response towards 0.1–1 ppm. (c) The repeatability, (d) selectivity, (e) interferent
selectivity and (f) long-term stability, respectively. Interfering gas mixtures in (e) include ethanol, acetone, toluene, formaldehyde, CO and H2, of
which the concentration applied is ten times as that of the target H2S.

gases were negligible. Additionally, we investigated the inter-


ferent selectivity in Fig. 4e, from which the responses of the H2S
mixture are observed to be close to those of the pure H2S,
further revealing the excellent selectivity of our BiFeO3/
Bi25FeO40 NPs sensors. Furthermore, we repeated the sensing
response in Fig. 4f on the 1st, 7th, 30th and 60th day under the
same evaluation conditions, and one can see that the sensing
responses are almost the same, exhibiting great stability.
Similarly, the sensing performance of the BiFeO3/Bi25FeO40
NPs toward the acetone biomarker under p-type sensing
behavior was investigated as well. The optimum working
temperature was found to be 300  C toward 1 ppm acetone in
Fig. 5a. Fig. 5b shows the real-time responses, and one can
observe that in p-type mode, acetone can be detected as low as
0.5 ppm. Furthermore, Fig. 5c shows the selectivity of the
BiFeO3/Bi25FeO40 NPs sensor to acetone at 300  C, and the
Fig. 5 Under p-type mode, the sensing performance of the inter- interfering gases are well excluded. Also, the interferent selec-
connected BiFeO3/Bi25FeO40 NPs sensor prototype toward acetone at tivity is investigated in Fig. S8a and b,† which further veries
300  C and RH of 30%. (a) The responses to 1 ppm acetone at various
that the sensor is capable of selectively detecting acetone.
temperatures. (b) The real-time response curve towards 0.5–10 ppm
acetone. The sensing evaluations correspond to (c) selectivity and (d) Additionally, the sensor presents excellent short-term and long-
short-term and long-term stability, respectively. term stability by circular testing toward 2 ppm acetone at
a certain time interval (Fig. 5d). Remarkably, our BiFeO3/
Bi25FeO40 NPs sensor exhibits high response to H2S and low
the responses and gas concentrations. Also, the repeated detection limit to both acetone and H2S biomarkers, compared
stability towards 5 ppm H2S is shown in Fig. 4c. The response with the unsatisfactory sensing performance of BiFeO3-based
and recovery time of theBiFeO3/Bi25FeO40 NPs sensor toward sensors (Table S1†). For practical application, we investigated
1 ppm H2S are 32 s and 68 s (Fig. S7†), respectively. In particular, the feasibility of breath gas analysis with the BiFeO3/Bi25FeO40
we purposely compared the response toward 10 ppm H2S, with NPs. In this way, the simulated halitosis and diabetes breath in
those toward 100 and 1000 ppm of other interfering gases, and Fig. 6a was evaluated by mixing 1 ppm of H2S and acetone
found that the responses in Fig. 4d towards the interfering inside the exhaled breath of healthy people.45 Accordingly, the

26008 | J. Mater. Chem. A, 2020, 8, 26004–26012 This journal is © The Royal Society of Chemistry 2020
View Article Online

Paper Journal of Materials Chemistry A

response to the simulated patient breath was lower than that of the band structure of the BiFeO3/Bi25FeO40 NPs. In Fig. S11a
the pure biomarker in Fig. 6b and c. Nevertheless, it was much and b,† the UPS valence band spectra show that the valence
higher than that of the healthy breath, so the patients can be band maximum (VBM) energy is around 2.67 eV.46 In Fig. S11c,†
distinguished from healthy ones. the UV-vis absorption spectrum shows an absorption band in
In particular, to avoid the interference of humidity, we the range of 200–400 nm, and the band gap value (Eg) is esti-
Published on 16 November 2020. Downloaded by University of California - Santa Barbara on 5/16/2021 1:30:12 PM.

simulated the relative humidity of exhaled breath, and the mated to be 3.16 eV by the plot of (ahn)1/2 versus (hn) in
response of the BiFeO3/Bi25FeO40 NPs under n-type and p-type Fig. S11d.†47 Correspondingly, the conduction band minimum
modes at a relative humidity of 55% and 85% was investi- (CBM) is 0.49 eV according to eqn (4). Also, according to the
gated as shown in Fig. S9,† indicating that our sensors are able Mott–Schottky analysis in Fig. 2b, the interconnected BiFeO3/
to work even when the humidity reaches 85%. Notably, we Bi25FeO40 NPs are an n-type semiconductor, so the Fermi level
further tested the sensor prototypes to simulate the case of (EF) marked with a red dashed line is above the intrinsic Fermi
working in critically wet conditions. Specically, we purposely level (Ei) (Fig. 7a).
mounted an intake grid covered with waterproof membrane,
which is able to block water, only allowing the ventilation of the E(CBM) ¼ E(VBM)  Eg (4)
target gases (e.g., H2S and acetone). In Fig. S10,† we sprayed
water over the “ventilating window” of the sensor prototype.
Even when the prototype was immersed in water for 120 The surface-adsorbed oxygen molecules were ionized into
minutes, it presented nearly the same sensing response aer oxygen species (O2, O, O2) by the thermally stimulated
properly wiping off the surface water and drying the waterproof processes of oxygen adsorption, dissociation, and charge
membrane (Fig. 6d and e), and one can verify the experimental transfer, which involve the electrons in the conducting band.
details in the ESI Video.† As a result, our sensors have been The reactions are presented in eqn (5)–(7).48–50
tested under critical conditions, and are able to resist the
O2(ads) + e / O2(ads) (5)
inuence of humidity, thus showing potential in future gas-
sensing detection of biomarkers via practical breath diagnoses. O2(ads) + e / 2O(ads) (6)

O(ads) + e / O2(ads) (7)


2.4 The sensing mechanism
The energy band of the BiFeO3/Bi25FeO40 NPs was used to In our cases, the critical temperature is 260–340  C, below
understand the n/p transition sensing mechanism. Accordingly, which the adsorption of oxygen species (O2, O) leads to the
the ultraviolet photoelectron spectroscopy (UPS) and ultraviolet- reduction of electrons and the formation of an electron deple-
visible (UV-vis) absorption spectra were obtained to determine tion layer (Fig. 7b). When reducing gas H2S reacts with O (eqn

Fig. 6 The investigation of the breath gas analysis of the BiFeO3/Bi25FeO40 NPs sensor prototypes. (a) Practical simulation of the exhaled breath
of a healthy experimenter, the breath was collected using a tedlar bag and taken out with a syringe. The responses of (b) simulated halitosis
breath, 1 ppm H2S and healthy breath and (c) simulated diabetes breath, 1 ppm acetone and healthy breath. (d) The real-time response plot of the
sensor covered with a waterproof membrane being immersed in deionized water for 120 min was compared with the same sensor (e) without the
waterproof membrane.

This journal is © The Royal Society of Chemistry 2020 J. Mater. Chem. A, 2020, 8, 26004–26012 | 26009
View Article Online

Journal of Materials Chemistry A Paper


Published on 16 November 2020. Downloaded by University of California - Santa Barbara on 5/16/2021 1:30:12 PM.

Fig. 7 Schematic energy-band variation. (a) Flat band situation with n-type conductivity, (b) oxygen adsorption leads to a depletion layer and n-
type surface conductivity or (c) an inversion layer (the EF lies below the Ei) and p-type surface conductivity. The sensing mechanism diagrams of
the interconnected BiFeO3/Bi25FeO40 NPs sensor prototype. (b1) N-type sensing mode and (c1) p-type sensing mode, respectively.

(8)), electrons are released back to the conduction band, nar- to the n/p transition sensing, and their separated band struc-
rowing the electron depletion layer, and thus presenting n-type tures of BiFeO3 and Bi25FeO40 can be found in the Band struc-
mode (Fig. 7b1). In contrast, in p-type mode, when the working ture section.† In short, the formation of the BiFeO3 and
temperature rises, strong adsorption of the oxygen atom (O2) Bi25FeO40 heterojunctions induces band-bending, which
captures more electrons, leading to the formation of an inver- promotes highly isolated thermal-stimulated carriers.28 In these
sion layer on the BiFeO3/Bi25FeO40 NPs surface,31,51 which is cases, the equilibrium of the separated electrons and holes is
marked with orange dashed boxes in Fig. 7c. In this case, the EF modulated by the temperature-dependent ionization reaction of
lies below the Ei, and the main carriers are holes on the BiFeO3/ the surface-adsorbing oxygen, and nally promotes the n/p
Bi25FeO40 NPs surface, resulting in p-type surface conductivity. transition sensing.
When acetone reacts with O2, holes are consumed (eqn (9)),
narrowing the hole accumulation layer (Fig. 7c1).
3. Conclusions
 
H2S(ads) + 3O (ads) / H2O + SO2 + 3e (8) In summary, a single sensor detecting dual H2S and acetone
biomarkers has been developed, via modulating the n/p semi-
CH3COCH3(ads) + 8O 2
(ads) + 16h / 3CO2 + 3H2O
+
(9)
conducting transition of interconnected BiFeO3/Bi25FeO40 NPs.
For the response of n- and p-type mixed modes marked with Meanwhile, the n/p semiconducting transition has been
red dashed circles in Fig. 3a–e, we interpret this as the transient experimentally observed. Remarkably, the BiFeO3/Bi25FeO40
change of majority carriers on the surface of the BiFeO3/ NPs sensor prototypes exhibit ppb-level detection limit and
Bi25FeO40 NPs. Specically, under p-type sensing, the majority excellent stability in both n- and p-type modes. Signicantly,
carriers of holes react with the high concentration reducing human breath sensing has been simulated for both H2S and
gases and are signicantly consumed, and nally the electrons acetone biomarkers, and provides an experimental basis for
become the majority carriers. In comparison, when the sensor practical application in early-diagnosis of halitosis and dia-
prototypes are exposed to air, strong adsorption of oxygen betes. Moreover, the sensors with a waterproof membrane were
atoms (O2) will form an inversion layer again. Additionally, it evaluated by critically immersing them in water for 120 min.
should be noted that the BiFeO3/Bi25FeO40 NPs may contribute The mechanism of this n/p type transition sensing is attributed
to the synergistic effect on temperature-dependent ionization of

26010 | J. Mater. Chem. A, 2020, 8, 26004–26012 This journal is © The Royal Society of Chemistry 2020
View Article Online

Paper Journal of Materials Chemistry A

adsorbed oxygen molecules and semiconducting band-bending 14 S.-Y. Jeong, J.-W. Yoon, T.-H. Kim, H.-M. Jeong, C.-S. Lee,
of BiFeO3/Bi25FeO40 NPs. The dual n/p semiconducting types Y. Chan Kang and J.-H. Lee, J. Mater. Chem. A, 2017, 5,
are modulated by the working temperature parameters, which 1446–1454.
enables the monitoring of two biomarkers, and thus this 15 E.-X. Chen, H.-R. Fu, R. Lin, Y.-X. Tan and J. Zhang, ACS Appl.
method could be employed to design miniaturized gas sensors Mater. Interfaces, 2014, 6, 22871–22875.
Published on 16 November 2020. Downloaded by University of California - Santa Barbara on 5/16/2021 1:30:12 PM.

and multi-disease diagnosis by means of human breath. Future 16 Y. H. Cho, Y. C. Kang and J.-H. Lee, Sens. Actuators, B, 2013,
theoretical investigations are expected on the n/p semi- 176, 971–977.
conducting transition of interconnected BiFeO3/Bi25FeO40 NPs. 17 M. Fleischer, S. Kornely, T. Weh, J. Frank and H. Meixner,
Sens. Actuators, B, 2000, 69, 205–210.
Conflicts of interest 18 J. W. Yoon, J. S. Kim, T. H. Kim, Y. J. Hong, Y. C. Kang and
J. H. Lee, Small, 2016, 12, 4159.
There are no conicts to declare. 19 G. Konvalina and H. Haick, ACS Appl. Mater. Interfaces, 2012,
4, 317–325.
Acknowledgements 20 V. V. Kondalkar, L. T. Duy, H. Seo and K. Lee, ACS Appl.
Mater. Interfaces, 2019, 11, 25891–25900.
This work was nancially supported by the National Natural 21 X. Xing, Z. Zhu, L. Du, D. Feng, J. Chen, S. Li and D. Yang,
Science Foundation of China (Grant No. 52072184), the Appl. Surf. Sci., 2020, 502, 144106.
Fundamental Research Funds for the Central Universities, 22 Y. Liu, X. Gao, F. Li, G. Lu, T. Zhang and N. Barsan, Sens.
Nankai University (Grant No. 63201179), and International Actuators, B, 2018, 260, 927–936.
Research Fellow of the Japan Society of the Promotion of 23 X.-Y. Xu and B. Yan, J. Mater. Chem. A, 2017, 5, 2215–2223.
Science (JSPS, Postdoctoral Fellowships for Research in Japan 24 D. Zhao, H. Huang, S. Chen, Z. Li, S. Li, M. Wang, H. Zhu and
(Standard), P18334). X. Chen, Nano Lett., 2019, 19, 3448–3456.
25 Y. Huang, W. Jiao, Z. Chu, G. Ding, M. Yan, X. Zhong and
References R. Wang, J. Mater. Chem. C, 2019, 7, 8616–8625.
26 Y.-Z. Chen, S.-W. Wang, C.-C. Yang, C.-H. Chung,
1 S. Nisreen, B. N. Gerald, F. Konrads, C. Silke, L. Marcis and Y.-C. Wang, S.-W. Huang Chen, C.-W. Chen, T.-Y. Su,
H. Hossam, Nano Lett., 2015, 15, 1288–1295. H.-N. Lin, H.-C. Kuo and Y.-L. Chueh, Nanoscale, 2019, 11,
2 O. Adams, M. Hakim, N. Shehada, Y. Y. Broza, S. Billan, 10410–10419.
R. Abdah-Bortnyak, A. Kuten, H. Haick, G. Peng and 27 Y.-C. Lee, Y.-L. Chueh, C.-H. Hsieh, M.-T. Chang, L.-J. Chou,
U. Tisch, Nat. Nanotechnol., 2009, 4, 669–673. Z. L. Wang, Y.-W. Lan, C.-D. Chen, H. Kurata and S. Isoda,
3 Y. Chen, Y. Zhang, F. Pan, J. Liu, K. Wang, C. Zhang, Small, 2007, 3, 1356–1361.
S. Cheng, L. G. Lu, Z. Zhang and X. Zhi, ACS Nano, 2016, 28 W. Li, X. Wu, J. Chen, Y. Gong, N. Han and Y. Chen, Sens.
10, 8169–8179. Actuators, B, 2017, 253, 144–155.
4 L. Wei, X. Lin, S. Kuang, X. Zhou, B. Dong, G. Lu and H. Song, 29 Y. Gong, X. Wu, X. Zhou, X. Li, N. Han and Y. Chen, Sens.
NPG Asia Mater., 2018, 10, 293–308. Actuators, B, 2019, 289, 114–123.
5 T. Konry and D. R. Walt, J. Am. Chem. Soc., 2009, 131, 13232– 30 H. Huang, H. Gong, C. L. Chow, J. Guo, T. J. White, M. S. Tse
13233. and O. K. Tan, Adv. Funct. Mater., 2011, 21, 2680–2686.
6 M. Phillips, K. Gleeson, J. M. Hughes, J. Greenberg, 31 A. Gurlo, N. Bârsan, A. Oprea, M. Sahm, T. Sahm and
R. N. Cataneo, L. Baker and W. P. Mcvay, Lancet, 1999, 353, U. Weimar, Appl. Phys. Lett., 2004, 85, 2280–2282.
1930–1933. 32 Z. Dai, C.-S. Lee, Y. Tian, I.-D. Kim and J.-H. Lee, J. Mater.
7 L. Pauling, A. B. Robinson, R. Teranishi and P. Cary, Proc. Chem. A, 2015, 3, 3372–3381.
Natl. Acad. Sci. U. S. A., 1971, 68, 2374–2376. 33 A. Gurlo, M. Sahm, A. Oprea, N. Barsan and U. Weimar, Sens.
8 H. J. O'Neill, S. M. Gordon, M. H. O'Neill, R. D. Gibbons and Actuators, B, 2004, 102, 291–298.
J. P. Szidon, Clin. Chem., 1988, 34, 1613–1618. 34 F. Gao, S. Dong, F. Yuan, T. Yu and J. M. Liu, Adv. Mater.,
9 M.-Y. Chuang, C.-C. Chen, H.-W. Zan, H.-F. Meng and 2007, 19, 2889–2892.
C.-J. Lu, ACS Sens., 2017, 2, 1788–1795. 35 T. Liu, Y. Zhang, X. Yang, X. Hao, X. Liang, F. Liu, F. Liu,
10 A. Kolmakov, D. O. Klenov, Y. Lilach, S. Stemmer and X. Yan, J. Ouyang and G. Lu, Sens. Actuators, B, 2018, 276,
M. Moskovits, Nano Lett., 2005, 5, 667–673. 489–498.
11 I.-S. Hwang, J.-K. Choi, H.-S. Woo, S.-J. Kim, S.-Y. Jung, 36 S. Das, S. Rana, S. M. Mursalin, P. Rana and A. Sen, Sens.
T.-Y. Seong, I.-D. Kim and J.-H. Lee, ACS Appl. Mater. Actuators, B, 2015, 218, 122–127.
Interfaces, 2011, 3, 3140–3145. 37 M. Dziubaniuk, R. Bujakiewicz-Korońska, J. Suchanicz,
12 J. M. Baik, M. Zielke, M. H. Kim, K. L. Turner, A. M. Wodtke J. Wyrwa and M. Re ˛kas, Sens. Actuators, B, 2013, 188, 957–
and M. Moskovits, ACS Nano, 2010, 4, 3117–3122. 964.
13 S. Jansat, K. Pelzer, J. Garcı́a-Antón, R. Raucoules, 38 V. A. Aksyuk, Nat. Nanotechnol., 2017, 12, 940–941.
K. Philippot, A. Maisonnat, B. Chaudret, Y. Guari, 39 W. Shi, Y. Guo and Y. Liu, Adv. Mater., 2020, 32, 1901493.
A. Mehdi, C. Reyé and R. J. P. Corriu, Adv. Funct. Mater., 40 Y. Chen, Q. Wu and J. Zhao, J. Alloys Compd., 2009, 487, 599–
2007, 17, 3339–3347. 604.

This journal is © The Royal Society of Chemistry 2020 J. Mater. Chem. A, 2020, 8, 26004–26012 | 26011
View Article Online

Journal of Materials Chemistry A Paper

41 N. B. Hakiki, S. Boudin, B. Rondot and M. D. C. Belo, Corros. 47 H. Tian, H. Fan, J. Ma, Z. Liu, L. Ma, S. Lei, J. Fang and
Sci., 1995, 37, 1809–1822. C. Long, J. Hazard. Mater., 2018, 341, 102–111.
42 Y. Liu, H. Guo, Y. Zhang, W. Tang, X. Cheng and W. Li, Chem. 48 W. Li, S. Ma, Y. Li, G. Yang, Y. Mao, J. Luo, D. Gengzang,
Eng. J., 2018, 343, 128–137. X. Xu and S. Yan, Sens. Actuators, B, 2015, 211, 392–402.
43 R. Köferstein, T. Buttlar and S. G. Ebbinghaus, J. Solid State 49 H. Yuan, S. A. A. A. Aljneibi, J. Yuan, Y. Wang, H. Liu, J. Fang,
Published on 16 November 2020. Downloaded by University of California - Santa Barbara on 5/16/2021 1:30:12 PM.

Chem., 2014, 217, 50–56. C. Tang, X. Yan, H. Cai, Y. Gu, S. J. Pennycook, J. Tao and
44 H. You, Z. Wu, L. Zhang, Y. Ying, Y. Liu, L. Fei, X. Chen, D. Zhao, Adv. Mater., 2019, 31, 1970076.
Y. Jia, Y. Wang, F. Wang, S. Ju, J. Qiao, C.-H. Lam and 50 Y. Zou, S. Chen, J. Sun, J. Liu, Y. Che, X. Liu, J. Zhang and
H. Huang, Angew. Chem., Int. Ed., 2019, 58, 11779–11784. D. Yang, ACS Sens., 2017, 2, 897–902.
45 S. J. Kim, S. J. Choi, J. S. Jang, H. J. Cho, W. T. Koo, 51 Y. C. Lee, Y. L. Chueh, C. H. Hsieh, M. T. Chang and S. Isoda,
H. L. Tuller and I. D. Kim, Adv. Mater., 2017, 29, 1700737. Small, 2010, 3, 1356–1361.
46 Y. Zhao, H. Fan, K. Fu, L. Ma, M. Li and J. Fang, Int. J.
Hydrogen Energy, 2016, 41, 16913–16926.

26012 | J. Mater. Chem. A, 2020, 8, 26004–26012 This journal is © The Royal Society of Chemistry 2020

You might also like