Next Article in Journal
Low-Power and Low-Cost Environmental IoT Electronic Nose Using Initial Action Period Measurements
Next Article in Special Issue
Development of Open-Tubular-Type Micro Gas Chromatography Column with Bump Structures
Previous Article in Journal
Monitoring the Land Subsidence Area in a Coastal Urban Area with InSAR and GNSS
Previous Article in Special Issue
A Binder Jet Printed, Stainless Steel Preconcentrator as an In-Line Injector of Volatile Organic Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sub-ppm Formaldehyde Detection by n-n TiO2@SnO2 Nanocomposites

1
Chemistry Department, Moscow State University, Moscow 119991 Russia
2
Faculty of Materials Science, Moscow State University, Moscow 119991 Russia
3
Department of Materials Science and Engineering, Inha University, Incheon 22212, Korea
4
School of Materials Science and Engineering, Hanyang University, Seoul 04763, Korea
*
Author to whom correspondence should be addressed.
Submission received: 12 June 2019 / Revised: 10 July 2019 / Accepted: 17 July 2019 / Published: 19 July 2019

Abstract

:
Formaldehyde (HCHO) is an important indicator of indoor air quality and one of the markers for detecting lung cancer. Both medical and air quality applications require the detection of formaldehyde in the sub-ppm range. Nanocomposites SnO2/TiO2 are promising candidates for HCHO detection, both in dark conditions and under UV illumination. Nanocomposites TiO2@SnO2 were synthesized by ALD method using nanocrystalline SnO2 powder as a substrate for TiO2 layer growth. The microstructure and composition of the samples were characterized by ICP-MS, TEM, XRD and Raman spectroscopy methods. The active surface sites were investigated using FTIR and TPR-H2 methods. The mechanism of formaldehyde oxidation on the surface of semiconductor oxides was studied by in situ DRIFTS method. The sensor properties of nanocrystalline SnO2 and TiO2@SnO2 nanocomposites toward formaldehyde (0.06–0.6 ppm) were studied by in situ electrical conductivity measurements in dark conditions and under periodic UV illumination at 50–300 °C. Nanocomposites TiO2@SnO2 exhibit a higher sensor signal than SnO2 and a decrease in the optimal measurement temperature by 50 °C. This result is explained based on the model considering the formation of n-n heterocontact at the SnO2/TiO2 interface. UV illumination leads to a decrease in sensor response compared with that obtained in dark conditions because of the photodesorption of oxygen involved in the oxidation of formaldehyde.

1. Introduction

Formaldehyde, HCHO, a colorless gas with an unpleasant odor, is a toxic compound that causes, in trace concentrations of 0.1–0.5 mg/m3, serious diseases of the respiratory tract, gastrointestinal tract and eyes. Biochemical oxidation of HCHO in human body occurs with the formation of carbon dioxide and formic acid, which, with prolonged exposure, causes asthma, pulmonary edema, and cancer. Formaldehyde is widely used in the manufacturing of polymeric materials for flooring, furniture, heat and electrical insulation, artificial tissues, plastic windows, etc. In addition, a concentrated HCHO solution (formalin) is used in medicine for disinfection, and in the food industry for the preservation of fruits and vegetables. Formaldehyde is also included in some cosmetics and personal care products. A detailed description of the characteristics of formaldehyde and its effects on health can be found in Ref. [1]. According to the World Health Organization, the maximum permissible concentration of formaldehyde in the air of the working area is 0.5 mg/m3 (0.4 ppm), in the air of the residential area is 0.125 mg/m3 (0.1 ppm) [1,2].
Recently, it was proposed that formaldehyde is one of the markers for lung cancer detection using exhaled breath analysis [3,4,5,6]. The median HCOH level observed for lung cancer patients (83 ppb) is higher than that for healthy ones (48 ppb) [5]. Thus, both medical and air quality applications require the detection of formaldehyde in the sub-ppm range.
Metal oxide based semiconductor gas sensor for formaldehyde detection are quite attractive since they can have sufficient sensitivity, are inexpensive and can be integrated into portable devices such as e-Nose, that could detect formaldehyde at ppb level at high humidity level (e.g., 90% RH for breath analysis) and in the presence of different interfering gases [6,7,8]. However, most studies consider the sensor characteristics of various semiconductor oxides in the detection of HCHO in the concentration range of 10 ppm or more [9,10]. Recent efforts are focused at the implementation of a new principle of HCHO detection under conditions of minimal thermal heating combined with illumination by a low-power UV or visible light source [11,12].
It was shown by the authors of [13,14,15,16,17,18] that under UV illumination molecules of volatile organic compounds (VOCs: acetone, acetaldehyde, ethanol, hydrocarbons) undergo photolysis on the surface of semiconductor oxides that facilitates their subsequent oxidation with chemisorbed oxygen, leading to a change in the conductivity of the semiconductor. Titanium dioxide TiO2 and zinc oxide ZnO demonstrate the highest activity in the photocatalytic VOCs oxidation. A comparative study on UV light activated porous TiO2 and ZnO sensors demonstrated that TiO2 exhibited a superior performance to ethanol and formaldehyde [19].
As a sensor material for HCHO detection TiO2 with different morphology was studied: mesoporous powders [20], nanotube arrays [21], polycrystalline thick films [22], hollow microspheres [23]. In all cases, the investigations were conducted under UV illumination. The experimental study of the TiO2 sensor properties in dark conditions is difficult because of the exceptionally high electrical resistance of these materials. A decrease in the resistance of TiO2-based materials can be achieved by creating nanocomposites, in which TiO2 plays the role of a photocatalyst/receptor, while the transport of charge carriers is effectuated by another semiconductor with more suitable electrical characteristics. Nanocomposites SnO2/TiO2 are promising candidates for formaldehyde detection both in dark conditions and under UV illumination. It was shown that as compared with pure SnO2 and TiO2, nanocomposites with different Sn/Ti ratio demonstrate higher sensor response toward triethylamine [24], ammonia [25], hydrogen [26,27], and formaldehyde [28]. Moreover, DFT calculations and experimental study presented by the authors of [29] revealed that these nanocomposites have higher photocatalytic activity in UV-light initiated degradation of RB5 and RhB dyes. However, the role of each component in the formation of the sensor signal when detecting formaldehyde in dark conditions and under UV illumination is not clear.
In this paper, we compare the sensor properties of nanocrystalline SnO2 and TiO2@SnO2 nanocomposites when detecting formaldehyde in the sub-ppm range in dark conditions and under periodic UV illumination in the temperature range of 50–300 °C.

2. Materials and Methods

2.1. Materials Synthesis

Nanocrystalline SnO2 was prepared by the precipitation method. Aqueous ammonia solution (25%) was added dropwise to 0.3 M solution of tin (IV) chloride pentahydrate (SnCl4·5H2O, 98%, Sigma-Aldrich) under vigorous stirring until pH = 7 [30,31]. The synthesis was carried out at room temperature. The obtained α-stannic acid gel was separated by centrifugation at 3500 rpm for 3 minutes, repeatedly washed with deionized water and then with 0.01 M NH4NO3 solution. AgNO3 test was used to confirm the negative reaction to chloride ions. The resulting gel was dried in air at 50 °C for 24 h, then obtained β-stannic acid xerogel was ground in an agate mortar and annealed in air at 300 °C for 24 h. According to transmission electron microscopy data [30,31] the SnO2 powder consists of a rounded shape, agglomerated nanoparticles of 3–8 nm in diameter.
Nanocomposites TiO2@SnO2 were synthesized by atomic layer deposition (ALD) method. Titanium (IV) isopropoxide (Ti(OCH(CH3)2)4, TTIP) and H2O used as precursors, were introduced consequently into the growth reactor containing SnO2 nanocrystals to avoid premature interaction. The ALD parameters: temperature and pressure were set to 150 °C and 0.1 torr, respectively. TTIP and H2O were kept in bubblers at 40 °C and 24 °C, respectively. The typical ALD cycle included the following stages: 1 s for TTIP dosing, 10 s for N2 purging, 0.2 s for H2O dosing, and 25 s for N2 purging. The ALD cycles were repeated for 500, 1000 and 2000 times, respectively, producing TiO2@SnO2 nanocomposites with increasing Ti/Sn ratio. After ALD synthesis the nanocomposites were annealed in air at 500 °C for 1 h to remove organic residues and ensure TiO2 crystallization. To obtain an adequate reference sample, nanocrystalline SnO2, which was not subjected to ALD treatment, was additionally annealed in air at 500 °C for 24 h.

2.2. Materials Characterization

Chemical composition ([Ti]/([Ti]+[Sn]), mol%) of TiO2@SnO2 nanocomposites was determined by X-ray fluorescent analysis (XRF) performed on M1 Mistral spectrometer (Bruker, Billerica, MA, USA) with the beam energy of 50 keV. The diameter of the analyzed area was 1.0 mm, the signal accumulation time 2 min.
The microstructure of the synthesized products was studied by high resolution transmission electron microscopy (FE-TEM, JEOL JEM-2100F, JEOL Ltd., Tokyo, Japan).
The phase composition of TiO2@SnO2 nanocomposites was characterized using powder X-ray diffraction (XRD) and Raman spectroscopy. The XRD patterns were collected by DRON-4 diffractometer (Burevestnik, Moscow, Russia) using monochromatic Cu Kα radiation (λ = 1.5406 Å). The survey was carried out in the range of 2θ = 10°–80° with a step of 0.1° at scanning rate of 0.5°/min. The crystallite size dXRD of SnO2 and TixOy phases was calculated from the broadening of the most intense XRD peaks using Scherer equation. WinXPow (STOE and Cie GmbH, Darmstadt, Germany) software was used for phase and full profile analysis, DSH software (laboratory made) was used for crystallite size calculations. The crystalline phases were determined using structural parameters of SnO2 cassiterite (ICDD 41-1445), TiO2 anatase (ICDD 21-1272) and TiO2 brookite (ICDD 75-1582). Raman spectra were recorded by a Renishaw InVia multichannel spectrometer (Renishaw plc., Gloucestershire, UK) using an argon laser with a wavelength of 514 nm as the radiation source. The laser beam was focused on the sample using a microscope objective (magnification × 50) so that the analyzed area was approximately 100 μm2. The spectra were recorded in air at room temperature in the range of 100–900 cm−1 with a step of 1 cm−1 (data was collected for each point for 10 s).
The nature of chemical groups presenting in obtained samples was investigated using the FT-IR spectroscopy. The IR spectra were recorded using a Perkin–Elmer Spectrum One Fourier Transform Infrared (FT-IR) spectrometer (Perkin Elmer Inc., Waltham, MA, USA) in the range of 4000–400 cm–1 with a step of 1 cm–1 in the transmission mode. For this the TiO2@SnO2 powders (0.3–0.5 mg) were grinded with 50 mg of dried (350 °C, 2 h) KBr (FT-IR grade, Sigma-Aldrich, Saint Louis, MO, USA) and pressed into tablets. DRIFT spectra were collected in situ using the DiffusIR annex and heated flow chamber HC900 (Pike Technologies, Fitchburg, WI, USA) sealed by ZnSe windows. DRIFT spectra were registered in the range 4000–1000 cm−1 at 4 cm−1 resolution with an accumulation of 30 scans. Powders (50 mg) were placed in alumina crucibles (5 mm diameter). Gas mixture containing 100 ppm of HCHO in dry air was used for the investigations.
The investigations of SnO2 reference sample and TiO2@SnO2 nanocomposites by the method of thermo-programmed reduction with hydrogen (TPR-H2) was effectuated on Chemisorb 2750 (Micromeritics, Norcross, GA, USA) in a quartz reactor using 10% H2/Ar gas mixture (50 mL/min) under the heating to 900 °C at a heating rate of 10 K/min.
Specially designed micro-hotplates were used to investigate gas sensor properties. The micro-hotplates consist of dielectric substrate (Al2O3) with dimensions of 0.9 × 0.9 × 0.15 mm, which provides a small temperature gradient between the heater and the sensitive layer. Substrate is covered with Pt electrodes on the top side for resistance measurements and Pt heater on the back side, which is protected by an insulating layer of dielectric paste. Electrodes and heater are made using Pt-based paste by screen printing method (Figure 1). The powders of nanocrystalline oxides were mixed with a vehicle (α-terpineol in ethanol) and deposited in the form of thick films over dielectric substrate to cover the electrodes. Thick films were sintered at 300 °C for 10 hours in air to remove the organic binder. The method used allows us to obtain the coatings which are continuous and uniform over the entire substrate with the thickness about 1 μm [30] and with the value of resistance in air at 300 °C, differing by no more than 10%. The current–voltage (I-V) characteristics of the sensors measured using Potentiostat P-8-NANO (Elins, Zelenograd, Russia) are shown in Figure 1c. All the samples exhibited linear I-V curves both for positive and negative applied bias voltages up to +2V and −2V, respectively.
The schematic illustration of sensor measurements setup is shown in Figure 2. The DC conductivity was measured in situ using electronic module (10) providing control of sensor heating and high-precision measurement of the resistance of the sensitive layer. The sensors were placed into a Teflon airtight and light-tight flow chamber (3) connected to a computer-controlled (5) gas delivery system with electronic mass-flow controllers (8). The measurements were carried out under a controlled constant flux of 100 ml/min in the temperature range of 300–50 °C with a step of 50 °C, in the dark conditions and under periodic UV illumination. Miniature UV LED (4) (10 mW/cm2, λmax = 365 nm) located at a distance of 4 cm above the sensors was used as an illumination source. The power of the LED was effectuated using DC power supply (1). The illumination of the sensors was carried out in a pulsed mode with the period 2 min “on”–2 min “off”. An automatic cyclic relay (2) periodically opened and closed the circuit, allowing measurements to be made with periodic illumination. The sensor properties of SnO2 reference sample and TiO2@SnO2 nanocomposites were investigated toward formaldehyde HCHO (0.06–0.08–0.15–0.3–0.6 ppm) in air (relative humidity at 25 °C RH25 = 30%). The gas mixtures containing preassigned concentration of HCHO were prepared by dilution of certified gas mixture (6) with background purified air (7). Flow through humidifier (9) was used to adjust the humidity (RH25 = 30%) of gases passing through the sensor chamber. The sensor response was calculated as S = R a i r / R g a s , where R a i r is the resistance in background air, R g a s is the resistance in the presence of HCHO.

3. Results and Discussion

3.1. Characteristics of Nanocrystalline SnO2 and TiO2@SnO2 Nanocomposites

The elemental and phase composition of the samples under investigation are presented in Table 1. An increase in the number of ALD cycles led to a proportional increase in the titanium content (presented as ([Ti]/([Ti]+[Sn]) ratio, mol%) in nanocomposites.
Figure 3 shows the TEM image of SnO2 powder used as a substrate in ALD synthesis confirming formation of SnO2 nanocrystals with approximate size of 5–10 nm. Lattice-resolved TEM image presented in Figure 3b highlights the crystalline phase of SnO2 with a plane spacing of 0.34 nm belonging to the (110) plane of cassiterite SnO2. The lattice-resolved TEM image of SnTi-1 nanocomposite (Figure 3c) makes it possible to detect crystalline regions with interplanar distances of 0.34 and 0.29 nm, corresponding to (110) SnO2 (cassiterite) and (121) TiO2 (brookite), respectively.
According to the results of the XRD analysis (Figure 4), the SnO2 annealed at 500 °C (equivalently to the post-synthetic annealing used for the synthesis of TiO2@SnO2 nanocomposites) crystallizes in a tetragonal cassiterite structure (ICDD 41-1445) with a crystallite size of 9–10 nm. Intense reflections from the cassiterite phase are observed in the diffraction patterns of all the samples. The titanium-containing phases, formed during post-synthetic annealing, change with an increase in the number of ALD cycles. The TiO2 with brookite structure (ICDD 29-1360) presents in SnTi-1 and SnTi-2 nanocomposites, as evidenced by the appearance of the intense (121) diffraction peak at 2θ = 30.8°. The most intense (120) at 2θ = 25.3° and (111) 2θ = 25.7° diffraction reflections of brookite coincide with the most intense (101) diffraction reflection (2θ = 25.3°) of the TiO2 anatase phase (ICDD 21-1272). Therefore, the presence of the anatase phase in these nanocomposites cannot be excluded. The intensity of diffraction peak at 2θ = 25.3° increases with increasing titanium content. In the diffraction pattern of SnTi-3 nanocomposite a new diffraction peak at 2θ = 48.1° corresponds to the (200) reflection of the anatase phase. At the same time, the diffraction peak at 2θ = 30.8° does not appear that indicates the absence of the brookite phase in this sample. The least intense and wide peaks in the ranges of 2θ = 36.5°–40.0° and 2θ = 52.6°–64.5° correspond to the superposition of the reflections of almost all of the above phases, so their deconvolution and assignment to a certain phase is a difficult task. The broadening of the SnO2 reflections in the diffraction patterns of nanocomposites and overlap of intense diffraction peaks corresponding to SnO2 and TiO2 in SnO2-TiO2 (anatase) and SnO2-TiO2 (brookite) systems (in contrast to the TiO2-ZnO [32]) do not allow us to make a positive or negative conclusion about the incorporation of Ti into the SnO2 lattice.
The crystallite size of detected Sn- and Ti-containing phases depending on titanium content in the samples is presented in Table 1. The crystallite size of tin dioxide, used as a substrate for the deposition of titanium dioxide by the ALD method, is 3–4 nm [30,31]. The post-synthetic annealing at 500 °C, necessary for the removal of organic residues of the titanium precursor, leads to an increase in the size of SnO2 crystallites to 9–10 nm. With an increase in the titanium content in TiO2@SnO2 nanocomposites, a decrease in the size of SnO2 crystal grains is observed as compared with the reference sample. A similar effect was repeatedly noted for nanocomposites based on semiconductor metal oxides [33,34,35]. The presence of impurities on the surface of growing crystallites slows down their growth rate under isothermal annealing due to the so-called Zener pinning [36]. The maximum crystallite size of the main phase is determined by the volume fraction and the size of particles (crystalline or amorphous) segregated on the surface of growing crystallites. The deposition of TiO2 layer on the surface of nanocrystalline SnO2 reduces the area of SnO2 intergranular contacts, that prevents recrystallization of SnO2 particles. The greater the thickness of the deposited TiO2 layer, the less the coarsening of SnO2 particles occurs. For the TiO2 brookite phase in SnTi-1 and SnTi-2 nanocomposites the crystallite size was estimated using the (121) reflection that does not overlap with another diffraction peaks. For the size of the crystallites of the TiO2 anatase phase, such an estimate is possible only in the case of SnTi-3 nanocomposite, in which there is no brookite phase. For the Ti-containing phases, the increase in the crystallite size is observed. Since during the ALD cycles the Ti-containing precursor has been deposited on the outer surface of SnO2 agglomerates, the sintering and coarsening of Ti-containing particles will be easier. That is why an increase in the titanium content leads to an increase in the crystallite size of the Ti-containing phases.
The IR spectra of nanocrystalline SnO2 and TiO2@SnO2 composites are compared in Figure 5a. In all spectra, a large broad peak in the region of 2700–3500 cm–1 is observed, which is related to the stretching vibrations of OH groups and a peak at 1628 cm–1, which is related to the deformation vibrations of adsorbed water [37,38]. The wide and intense absorption bands at 530 cm–1 and 620 cm–1 in the case of SnO2 correspond to stretching vibrations of the Sn–O and symmetric vibrations of the O–Sn–O bonds, respectively [39]. The metal–oxygen (M–O) oscillation modes for SnO2 and TiO2 in TiO2@SnO2 composites overlap in the range of 400–650 cm–1, nevertheless different components can be distinguished. All the FT-IR spectra of the composites show vibration modes at 450 and 610 cm–1, which can be attributed to oscillations of Ti–O–Ti and Ti–O bonds, respectively [40,41,42]. With an increase in the titanium content an increase in the intensity of the Ti–O oscillation modes is observed, which appears as a broadening of the M–O absorption peak in this region.
Figure 5b shows the Raman spectra of nanocrystalline SnO2 (SnTi-0) and TiO2@SnO2 composites. Raman spectrum of the SnO2 clearly shows three characteristic modes at 480.5, 631 and 772.8 cm−1 that correspond to the Eg, A1g and B2g vibrational modes, respectively [43,44]. The A1g and B2g modes are associated with symmetric and asymmetric Sn–O stretching, respectively, orthogonally to the c axis. The translational Eg mode is related to the motion of oxygen anions along the c axis [43,44]. The B1g oscillation mode appears only in the spectra of nanocrystalline SnO2 in the range of 100–184 cm−1 [45,46]. In the SnO2 spectrum presented in Figure 5b the band at 137 cm−1 is due to B1g mode, the shift may be associated with the nanoparticles size effect.
J. Zuo et al. studied the size effects in SnO2 nanoparticles [47]. They showed that, in addition to the characteristic vibrational modes of bulk SnO2, the Raman spectrum of nanocrystalline SnO2 has two additional Raman scattering bands at 358 (B1) and 572 cm−1 (B2). The B2 band corresponds to surface modes and is very sensitive to the changes in crystallite size for nanoscale SnO2. The appearance of surface modes is associated with a small particle size of SnO2 and can be due to the appearance of oscillations that are forbidden by symmetry due to the breaking of long-range order in the systems of reduced dimension. That is why a decrease in crystallite size leads to the formation of a highly defective surface layer, the contribution of which will be the highest for the materials with the smallest particle size [48]. Based on the above considerations, the wide band located at 563 cm−1 corresponds to the surface modes associated with in-plane oxygen vacancies of the nanocrystalline cassiterite SnO2 [48,49,50]. Consequently, the small size and surface defects of the SnO2 nanoparticles may have a positive effect on the gas sensitivity of the sensor. This assumption was experimentally confirmed in by the authors of [31], where it was shown that the relative intensity of Raman surface modes IS/IV taken as the ratio of the sum of their intensities IS to the intensity of A1g mode IV demonstrates the best linear correlation with gas response of SnO2 nanocrystalline materials to CO.
Changes in the crystal’s local symmetry produce changes in some of the components of the polarizability tensor, even for usually forbidden vibration modes [49]. That is why the A2u IR active and Raman forbidden modes are found to transform into Raman active modes [48]. In this case the bands at 309 and 352 cm−1 (Eu) and the band at 444 (B1u) are related to transformation of an IR to Raman active modes.
Raman peaks observed in the spectra of TiO2@SnO2 composites at 145, 198, 397, 516 and 635 cm−1 (Figure 5b), refer to the Eg, Eg, B1g, A1g + B1g and Eg modes of the anatase phase, respectively [51]. For TiO2 nanoparticles the Eg Raman peak is mainly caused by symmetric stretching vibration of O–Ti–O groups, B1g peak is caused by symmetric bending vibration of O–Ti–O and A1g peak is caused by antisymmetric bending vibration of O–Ti–O [52]. The presence of an intense 145 cm−1 mode (a characteristic oscillation mode of anatase) indicates that TiO2 nanocrystals have a certain degree of long-range order, while weaker and wider peaks in the high-frequency region indicate the absence of a short-range order in the anatase phase [53,54,55]. According to the factor group analysis, TiO2 brookite phase has a total of 36 Raman active modes (9A1g + 9B1g + 9B2g + 9B3g). Raman spectra of the SnTi-1 and SnTi-2 composites show both anatase and brookite bands. In total, 4 brookite bands were readily identified, including A1g (255 cm−1), B3g (285, 443 cm−1), B2g (581 cm−1). Also, the A1g (153, 194 cm−1) and B2g (395 cm−1) brookite bands may be overlapped by anatase modes, which are very close to them [56,57]. The full assignment of the vibrational modes in the Raman spectra of nanocomposites is presented in Table 2.
Thus, the Raman spectra confirm the data obtained by the XRD method. The formation of the metastable brookite phase is observed in SnTi-1 and SnTi-2 composites. Anatase can play the role of a stabilizer for this phase [58]. A correlation among the surface enthalpies of the TiO2 three polymorphs and their particle size was found by Zhang and Banfield [59]. The formation energies of anatase, brookite and rutile are sufficiently close that they can be reversed by small differences in surface energies. Zhu et al. [60] developed an empirical expression on a critical grain size of brookite, which determines the transition sequence between anatase and brookite. These transformations become noticeable with prolonged isothermal annealing at temperatures above 500 °C. To avoid the changes in phase composition and crystallite size of nanocomposites subjected to post synthesis annealing at 500 °C, the maximum temperature during the manufacture of sensor elements and gas sensor measurements did not exceed 300 °C.

3.2. Gas Sensor Properties

As discussed in the review [8], depending on the operating temperature the change of the resistance of n-type semiconductor oxides when interacting with formaldehyde is due to HCHO oxidation with chemisorbed oxygen O β ( a d s ) α to HCOOH or CO2:
β · H C H O ( g a s ) + O β ( a d s ) α = β · H C O O H ( g a s ) + α · e
β · H C H O ( g a s ) + 2 O β ( a d s ) α = β · C O 2 ( g a s ) + β · H 2 O ( g a s ) + 2 α · e
At constant temperature and HCHO concentration, the value of the sensor response will depend on the concentration and the predominant form of chemisorbed oxygen on the surface of the semiconductor oxide.
According to the literature [20,61,62,63,64], the interaction of UV light with TiO2 involves the following processes: electron-hole pairs generation (3), oxygen photodesorption (4), formation of “active” chemisorbed oxygen (5), which is able to oxidize formaldehyde even at room temperature (6)
h ν T i O 2 h + ( h ν ) + e ( h ν )
h + ( h ν ) + O 2 ( a d s ) O 2 ( g a s )
O 2 ( g a s ) + e ( h ν ) O 2 ( a d s ) ( h ν )
H C H O ( g a s ) + O 2 ( a d s ) ( h ν ) C O 2 ( g a s ) + H 2 O ( g a s ) + e
Under UV light, hydroxyl groups presented on the TiO2 surface can also pass into the “active” form (7) and then participate in the oxidation of formaldehyde (8):
O H + h + ( h ν ) O H ( h ν )
H C H O ( g a s ) + 2 · O H ( h ν ) C O 2 ( g a s ) + H 2 O ( g a s ) + 2 · H + + 2 · e
Figure 6 demonstrates the change in the resistance of nanocomposites with the cyclic changes in the composition of the gas phase “air (15 min)”–“0.6 ppm HCHO in air (15 min)”. The measurements were carried out in the temperature range of 300–50 °C in dark conditions (Figure 6a) and under periodic UV illumination (Figure 6b). At a fixed temperature, TiO2@SnO2 nanocomposites have a higher resistance than unmodified SnO2 (SnTi-0). UV illumination leads to reduced resistance of nanocomposites by 1.5–3 times. In all cases in the presence of HCHO, the resistance of SnO2 and TiO2@SnO2 nanocomposites decreases due to the oxidation of formaldehyde by chemisorbed oxygen. In dark conditions at low measurement temperatures (T < 200 °C), the baseline resistance drift is observed for all samples. This may be due to the accumulation of formaldehyde oxidation products on the surface of the sensitive layer under these conditions. The use of UV illumination reduces the drift of the resistance at T = 150 °C. It can be assumed that UV light stimulates the desorption of the products of formaldehyde oxidation.
The obtained results allowed us to draw the temperature dependence of the sensor response S = R a i r / R g a s (Figure 7a). The maximum sensor response of unmodified SnO2 (SnTi-0) is observed at T = 20 °C. The introduction of Ti-containing phases leads to a decrease in the temperature of the maximum sensor signal to 150 °C. In dark conditions, the maximum response was detected for the SnTi-2 nanocomposite. The use of UV illumination does not change the position of the maxima on the temperature dependence of the sensor response, but unexpectedly it leads to a slight decrease in the signal value in the low temperature range of 50–150 °C. Such a decrease in the sensor response can be caused by a decrease in the concentration of chemisorbed oxygen participating in the oxidation of formaldehyde by reactions (1) and (2), due to the partial desorption of oxygen from the surface of semiconductor oxides under UV light [65]. Figure 7b shows the temperature dependences of the effective photoresponse SPh of nanocomposites in background air calculated as SPh = Rdark/Rlight, where Rlight is the minimum resistance achieved during the sensor illumination, and Rdark is the maximum resistance achieved in the dark period [66]. The maximum photoresponse corresponds to a temperature range of 150–200 °C. A decrease in the SPh value with an increase in the measurement temperature up to 250–300 °C is due to the contribution of thermal oxygen desorption in dark conditions.
The dynamic sensor characteristics, response time τ 90 . and recovery time τ 90 * . , are presented in Figure 8. Even though the absolute values of response time and recovery time are strongly dependent on the parameters of the testing system (geometry and size of the measurement cell, the method to switch the gases, the gas flow rate), they are useful to compare the dynamic sensor characteristics of materials if measurements are performed in identical conditions. With a decrease in the operating temperature, an increase in both the response and recovery times is observed. It should be noted that the sensor based on unmodified SnO2 is characterized by significantly worse dynamic characteristics than sensors based on TiO2@SnO2 nanocomposites, which are close to each other. Such a difference may be due to a decrease in the degree of sintering of SnO2 nanoparticles in nanocomposites as compared to unmodified tin dioxide.
To clarify the mechanism of formaldehyde oxidation on the surface of semiconductor oxides, the in situ DRIFT studies have been conducted. In situ DRIFT spectra of the unmodified SnO2 (SnTi-0) and SnTi-2nanocomposite during HCHO adsorption at room temperature are shown in Figure 9.
The bands of weakly bonded forms at 1202, 1742, 1766 and 3010 cm−1 for SnTi-0 and at 1768 cm−1 for SnTi-2 attributed to molecular adsorption form of HCHO with the formation of hydrogen bonds between its carbonyl oxygen and surface hydroxyl groups [67,68,69]. The formation of CO2 (at 2337 and 2365 cm−1) was observed on SnTi-0 after 5 min of HCHO adsorption, moreover these bands grew in intensity with the increase of the adsorption time [70]. In SnTi-2 spectra the bands at 1344 and 1592 cm−1 are assigned to COO symmetric stretching and COO asymmetric stretching of formate species [71,72]. The bands located at 2838, 2890 and 2590, 2868 cm−1 for SnTi-0 and SnTi-2, respectively, belong to CH stretching of formate species [68,73,74]. The bands at 2936 cm−1 for SnTi-0 and 2930 cm−1 for SnTi-2 were assigned to the characteristic peak of dioxymethylene (DOM) intermediate [71]. The other DOM bands also were identified at 1050, 1107, 1145 cm−1 for SnTi-0 and 1066, 1107, 1146, 1165, 1478 cm−1 for SnTi-2 [67,68].
Figure 10 shows the DRIFT spectra of the SnTi-2 nanocomposite obtained during the heating in dry air at T = 300 °C after formaldehyde adsorption at room temperature. From the results, it can be seen that the formats are intensively desorbed from the surface of the nanocomposite, as evidenced by decreasing of their characteristic vibrational modes at 1346, 1382 and 1553 cm−1 [68,71,72]. However, at such a high temperature, there is an increase in the intensity of dioxymethylene species, the bands of which are at 1050, 1108 and 1450 cm−1 [67,68], which indicates partially HCHO oxidation to DOM. Most likely, at T=300 °C, formaldehyde is oxidized to CO2 and H2O, as a result of which hydroxyl groups of water accumulate on the surface (increase in the intensity of OH bands at 3550 and 3630 cm−1) and CO2 desorption occurs (decrease in intensity at 2340 cm−1); at room temperature CO2 is accumulated on the SnTi-0 surface (Figure 9a). The assignment of IR bands for intermediates is summarized in Table 3.
So, for HCHO oxidation, oxygen ions adsorbed on the surface of the nanocomposites play a key role in the generation of the surface formates. According to the results of the in situ DRIFT analysis and literature review [67,68,69,70,75,76] we propose a reaction mechanism as depicted in (9)–(12):
H C H O ( g a s ) H C H O ( a d s )
β · H C H O ( a d s ) + O β ( a d s ) α β · H 2 C O O ( a d s ) + ( α 1 ) · e
β · H 2 C O O ( a d s ) + O β ( a d s ) α β · H C O O ( a d s ) 2 + β · O H ( a d s ) ( α 1 )
2 β · H C O O ( a d s ) 2 + O β ( a d s ) α 2 β · C O 2 + β · H 2 O + ( 4 β + α ) · e
At first, formaldehyde molecularly adsorbs as a whole via strong hydrogen-bonding interactions (9). Desorption of formaldehyde is in competition with the reaction between formaldehyde and co-adsorbed surface oxygen to yield formate HCOO-, which can proceed via formation of dioxymethylene H2COO- intermediate (10),(11) [69]. For SnTi-0 sample there are two weak bands at 2337 and 2365 cm−1 indicated the appearance of carbon dioxide [64,77]. In this case, CO2 and H2O were formed as the result of oxidation of formate ions (12) of chemisorbed oxygen on semiconductor oxides [78], reaction (10) proceeds without changing the electron concentration in the conduction band of the semiconductor, that causes the absence of a sensor response at T = 50 °C (Figure 6).
The observed effect of Ti-containing phases on the sensor response of SnO2 toward HCHO in dark conditions should be explained based on the model considering the formation of n-n heterocontact at the SnO2/TiO2 interface. The estimated band alignment of the SnO2 (cassiterite) and TiO2 (anatase) is presented on Figure 11a [24,79,80]. The available lower-energy conduction band states stimulate electron transfer to n-SnO2 [81]. As a result, the depletion layer is formed at n-TiO2 surface due to loss of electrons, and accumulation layer is formed at SnO2 surface due to added electrons. In turn, an increase in the electron concentration stimulates the chemisorption of oxygen species enhancing the response formed due to reactions (1). The scheme in Figure 11a is constructed for the heterocontact SnO2 (cassiterite)/TiO2 (anatase), since the anatase phase is present in all nanocomposites (as it follows from the data of Raman spectroscopy). Brookite (revealed in SnTi-1 and SnTi-2 nanocomposites) is the least studied polymorph of TiO2. However, from the data of Ref. [82] it follows that the bottom of the conduction band of brookite lies above the bottom of the conduction band of anatase. Thus, in the formation of a heterocontact SnO2 (cassiterite)/TiO2 (brookite), electron transfer will also occur from TiO2 (brookite) to SnO2, increasing the chemisorption of oxygen on its surface. An increase in the concentration of chemisorbed oxygen on the surface of nanocomposites compared with unmodified SnO2 is confirmed by the method of thermo-programmed reduction with hydrogen (TPR-H2). Figure 11b shows the temperature dependencies of the hydrogen consumption during the reduction of unmodified SnO2 and SnTi-1 nanocomposite. The reduction of SnO2 with hydrogen occurs in the temperature range of 430–800 °C, the maximum H2 consumption is observed at T = 620 °C. The introduction of Ti-containing phases leads to a shift of the H2 consumption towards lower temperatures 570–600 °C. Such a change in the TPR-H2 profiles may be due to the reduction of highly dispersed Ti-containing phases. According to the reports [83,84,85], the reduction of TiO2 with hydrogen occurs at T > 400 °C, the maximum H2 consumption is observed at T = 530–650 °C, depending on the TiO2 dispersion. Another feature of the TPR-H2 profiles is the presence of hydrogen absorption peaks in the low-temperature range of 100–400 °C. This corresponds to the interaction of hydrogen with oxygen-containing particles (chemisorbed oxygen, hydroxyl groups, etc.) on the surface of highly dispersed semiconductor oxides. The reduction of TiO2@SnO2 nanocomposites is accompanied by more significant hydrogen consumption in the low-temperature region that indicates a higher concentration of oxygen-containing species on its surface. An increase in the concentration of chemisorbed oxygen on the surface of nanocomposites provides an increase in their resistance compared to unmodified SnO2. At the same time, due to the photodesorption of chemisorbed oxygen under UV light, TiO2@SnO2 nanocomposites are characterized by the larger photoresponse in the background air and demonstrate a more significant decrease in the sensor response to HCHO.
Thus, we are forced to conclude that in the case of TiO2@SnO2 nanocomposites, the amplitude of the sensor response when detecting HCHO in the sub-ppm range is determined mainly by the SnO2/TiO2 interface. The role of UV light is to enhance the photodesorption of oxygen, and the processes of HCHO oxidation by photoactivated particles (reactions (6) and (8)) do not contribute to the sensor response.
The dependence of the sensor response on the HCHO concentration in air was studied in dark conditions and under periodic UV illumination at 150 °C during the stepwise increase and subsequent stepwise decrease in the HCHO concentration (Figure 12). The results obtained allowed us to build the calibration curves (Figure 13a) and to determine the lower detection limit (LDL). As in the case of measurements at different temperatures, when using UV illumination, a decrease in the sensor response toward HCHO is observed, that may be due to the photodesorption of chemisorbed oxygen under UV light. The LDL values were calculated from the calibration curves as the gas concentration corresponding to the minimum measurable sensor signal Rav/(Rav – 3σ) , where Rav is the average sensor resistance in air at 150 °C, σ is the standard deviation of the resistance in air. A decrease in the sensitivity of nanocomposites is accompanied by an increase in LDL from 9–15 to 30–32 ppb (Table 4).
The comparison of the sensor response toward 0.06–0.6 ppm HCHO for SnTi-2 nanocomposite with the literature data [7,8,9,10,11,20,21,22,23,28,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117] is shown in Figure 13b and Table 5. There are very few papers that consider the detection of formaldehyde in a practically important sub-ppm concentration range. TiO2@SnO2 nanocomposites exhibit high sensitivity to HCHO in the sub-ppm range. Higher sensor response values were obtained only by the authors of [86], where metastable In4Sn3O12 was used as sensitive material at 350 °C and in Ref. [7], where Si modified SnO2 was used at 400 °C. It should be noted that TiO2@SnO2 nanocomposites have comparable sensitivity at much lower measurement temperature (150 °C vs. 350 °C or 400 °C), which can provide a significant reduction in the power consumption of semiconductor gas sensor.
In addition to reducing power consumption, a sufficiently low measurement temperature allows one to achieve an increase in the selectivity of sensors when detecting various VOCs. The cross sensitivity of SnO2 and TiO2@SnO2 nanocomposites was studied at a measurement temperature of 150 °C in the detection of formaldehyde, benzene, and acetone (Figure 14). It can be noted that the synthesized materials have low cross-sensitivity to benzene in the wide concentration range. As for acetone, the interference of signals can occur if acetone concentration is in the range of C > 1 ppm and at least 20 times higher than formaldehyde concentration. SnTi-2 nanocomposite is characterized by the lowest response to acetone.
Comparing the sensor characteristics of unmodified SnO2 and TiO2@SnO2 nanocomposites in the detection of formaldehyde at 150 °C (Table 4) we can conclude that the SnO2 is inferior in the magnitude of the sensor response and the dynamic characteristics (response and recovery time). For HCHO detection at concentrations of 60 ppb and above, the SnTi-2 nanocomposite, which exhibits the highest sensor response and the lowest cross-sensitivity to acetone, seems to be preferred. However, for the measurements of lower HCHO concentrations, the SnTi-1 nanocomposite characterized by lower LDL value, lower base resistance in air, and the same dynamic characteristics, may be optimal.
As a summary, we have to conclude, that compared to the sensor characteristics described in [7,86,118], the sensors based on TiO2@SnO2 nanocomposites are inferior in sensitivity and selectivity. However, their advantage is significantly reduced operating temperature. The increase in sensitivity while maintaining a low operating temperature should be possible due to the addition of modifiers of different chemical nature on the surface of TiO2@SnO2 nanocomposites. The increase in selectivity, in particular towards humidity, can be achieved using passive filters—selective membranes based on SiO2 [6,118,119,120] or metal organic frameworks [110,121] as well as by the creation of multi-sensor systems operating using mathematical signal processing [6,122].

4. Conclusions

Nanocomposites TiO2@SnO2 obtained by ALD synthesis were investigated as sensitive materials for sub-ppm formaldehyde detection in dark conditions and under periodic UV (λmax = 365 nm) illumination. As observed by XRD and Raman spectroscopy, all nanocomposites contain nanocrystalline SnO2 cassiterite and TiO2 anatase as the main crystalline phases. Depending on Ti content in nanocomposites predetermined by the number of ALD cycles, the minor TiO2 phases brookite, rutile and anosovite have been found. A thorough study of nanocomposites using FTIR and TPR-H2 methods made it possible to establish that nanocomposites have a higher concentration of chemisorbed oxygen and surface hydroxyl groups compared to unmodified SnO2. When detecting formaldehyde in the sub-ppm range, TiO2@SnO2 nanocomposites exhibit a higher sensor signal than SnO2 and a decrease in the optimal measurement temperature by 50 °C. This result is explained based on the model considering the formation of n-n heterocontact at the SnO2/TiO2 interface. TiO2@SnO2 nanocomposites have high sensitivity toward HCHO at quite low measurement temperature 150 °C that can provide a significant reduction in the power consumption and an increase in the selectivity of semiconductor gas sensors when detecting various VOCs.
Unexpectedly, UV illumination leads to a decrease in sensor response compared with the results obtained in dark conditions. So, we have to conclude that in the case of TiO2@SnO2 nanocomposites, the amplitude of the sensor response toward sub-ppm HCHO concentrations is determined mainly by the SnO2/TiO2 interface. The UV light stimulates the photodesorption of oxygen, while the processes of HCHO oxidation by photoactivated particles do not contribute to the sensor response.

Author Contributions

Conceptualization, M.R., A.G., S.S.K., H.W.K.; methodology, M.R., A.M., S.S.K.; formal analysis, A.N.; investigation, J.-H.L., J.-H.K., J.-Y.K., A.N., A.M.; data curation, A.N.; writing—original draft preparation, A.N., M.R.; writing—review and editing, M.R., A.G., S.S.K., H.W.K.; supervision, M.R.

Funding

The work was performed within the framework of the project of the Joint Research Program between Korea and Russia. Korean partner research was funded by the Ministry of Science and ICT of Korea (NRF-2017K1A3A1A49070046) under the International Research and Development Program of the National Research Foundation of Korea (NRF). The research of the Russian partner was funded by the Russian Ministry of Education and Science (Agreement No. 14.613.21.0075, RFMEFI61317X0075).

Acknowledgments

The work was performed using the equipment of the Center for Collective Use of Moscow State University “Technologies for the production of new nanostructured materials and their comprehensive research”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Health Organization (WHO). WHO Guidelines for Indoor Air Quality: Selected Pollutants. Available online: www.euro.who.int_data/assets/pdf_file/0009/128169/e94535/pdf (accessed on 7 July 2019).
  2. Nielsen, G.D.; Larsen, S.T.; Wolkoff, P. Re-evaluation of the WHO (2010) formaldehyde indoor air quality guideline for cancer risk assessment. Arch. Toxicol. 2017, 91, 35–61. [Google Scholar] [CrossRef]
  3. Konvalina, G.; Haik, H. Sensors for Breath Testing: From Nanomaterials to Comprehensive Disease Detection. Acc. Chem. Res. 2014, 47, 66–76. [Google Scholar] [CrossRef] [PubMed]
  4. Hakim, M.; Broza, Y.Y.; Barash, O.; Peled, N.; Phillips, M.; Amann, A.; Haick, H. Volatile Organic Compounds of Lung Cancer and Possible Biochemical Pathways. Chem. Rev. 2012, 112, 5949–5966. [Google Scholar] [CrossRef] [PubMed]
  5. Fuchs, P.; Loeseken, C.; Schubert, J.K.; Miekisch, W. Breath gas aldehydes as biomarkers of lung cancer. Int. J. Cancer 2010, 126, 2663–2670. [Google Scholar] [CrossRef] [PubMed]
  6. Güntner, A.T.; Abegg, S.; Königstein, K.; Gerber, P.A.; Schmidt-Trucksäss, A.; Pratsinis, S.E. Breath Sensors for Health Monitoring. ACS Sens. 2019, 42, 268–280. [Google Scholar] [CrossRef] [PubMed]
  7. Gütner, A.T.; Koren, V.; Chikkadi, K.; Righettoni, M.; Pratsinis, S.E. E-Nose Sensing of Low-ppb Formaldehyde in Gas Mixtures at High Relative Humidity for Breath Screening of Lung Cancer? ACS Sens. 2016, 1, 528–535. [Google Scholar] [CrossRef]
  8. Lv, P.; Tang, Z.; Wei, G.; Yu, J.; Huang, Z. Recognizing indoor formaldehyde in binary gas mixtures with a micro gas sensor array and a neural network. Meas. Sci. Technol. 2007, 18, 2997–3004. [Google Scholar] [CrossRef]
  9. Castro-Hurtado, I.; Mandayo, G.G.; Castaño, E. Conductometric formaldehyde gas sensors. A review: From conventional films to nanostructured materials. Thin Solid Films 2013, 548, 665–676. [Google Scholar] [CrossRef]
  10. Mirzaei, A.; Leonardi, S.G.; Neri, G. Detection of hazardous volatile organic compounds (VOCs) by metal oxide nanostructures-based gas sensors: A review. Ceram. Int. 2016, 42, 15119–15141. [Google Scholar] [CrossRef]
  11. Espid, E.; Taghipour, F. UV-LED Photo-activated Chemical Gas Sensors: A Review. Crit. Rev. Solid State Mater. Sci. 2017, 42, 416–432. [Google Scholar] [CrossRef]
  12. Xu, F.; HO, H.P. Light-Activated Matel Oxide Gas Sensors: A Review. Micromachines 2017, 8, 333. [Google Scholar] [CrossRef] [PubMed]
  13. Trawka, M.; Smulko, J.; Hasse, L.; Granqvist, C.-G.; Annanouch, F.E.; Ionescu, R. Fluctuation enhanced gas sensing with WO3-based nanoparticle gas sensors modulated by UV light at selected wavelengths. Sens. Actuators B 2016, 234, 453–461. [Google Scholar] [CrossRef]
  14. Wongrat, E.; Chanlek, N.; Chueaiarrom, C.; Samransuksamer, B.; Hongsith, N.; Choopun, S. Low temperature ethanol response enhancement of ZnO nanostructures sensor decorated with gold nanoparticles exposed to UV illumination. Sens. Actuators A 2016, 251, 188–197. [Google Scholar] [CrossRef]
  15. Cui, J.; Wang, D.; Xie, T.; Lin, Y. Study on photoelectric gas-sensing property and photogenerated carrier behavior of Ag–ZnO at the room temperature. Sens. Actuators B 2013, 186, 165–171. [Google Scholar] [CrossRef]
  16. Peng, L.; Zhao, Q.; Wang, D.; Zhai, J.; Wang, P.; Pang, S.; Xie, T. Ultraviolet-assisted gas sensing: A potential formaldehyde detection approach at room temperature based on zinc oxide nanorods. Sens. Actuators B 2009, 136, 80–85. [Google Scholar] [CrossRef]
  17. Peng, L.; Xie, T.; Yang, M.; Wang, P.; Xu, D.; Pang, S.; Wang, D. Light induced enhancing gas sensitivity of copper-doped zinc oxide at room temperature. Sens. Actuators B 2008, 131, 660–664. [Google Scholar] [CrossRef]
  18. De Lacy Costello, B.P.J.; Ewen, R.J.; Ratcliffe, N.M.; Richards, M. Highly sensitive room temperature sensors based on the UV-LED activation of zinc oxide nanoparticles. Sens. Actuators B 2008, 134, 945–952. [Google Scholar] [CrossRef]
  19. Chen, H.; Liu, Y.; Xie, C.; Wu, J.; Zeng, D.; Liao, Y. A comparative study on UV light activated porous TiO2 and ZnO film sensors for gas sensing at room temperature. Ceram. Int. 2012, 38, 503–509. [Google Scholar] [CrossRef]
  20. Liu, L.; Li, X.; Dutta, P.K.; Wang, J. Room temperature impedance spectroscopy-based sensing of formaldehyde with porous TiO2 under UV illumination. Sens. Actuators B 2013, 185, 1–9. [Google Scholar] [CrossRef]
  21. Lin, S.; Li, D.; Wu, J.; Li, X.; Akbar, S.A. A selective room temperature formaldehyde gas sensor using TiO2 nanotube arrays. Sens. Actuators B 2011, 156, 505–509. [Google Scholar] [CrossRef]
  22. Zhang, S.; Lei, T.; Li, D.; Zhang, G.; Xie, C. UV light activation of TiO2 for sensing formaldehyde: How to be sensitive, recovering fast, and humidity less sensitive. Sens. Actuators B 2014, 202, 964–970. [Google Scholar] [CrossRef]
  23. Li, X.; Li, X.; Wang, J.; Lin, S. Highly sensitive and selective room-temperature formaldehyde sensors using hollow TiO2 microspheres. Sens. Actuators B 2015, 219, 158–163. [Google Scholar] [CrossRef]
  24. Jia, C.; Dong, T.; Li, M.; Wang, P.; Yang, P. Preparation of anatase/rutile TiO2/SnO2 hollow heterostructures for gas sensor. J. Alloy. Compd. 2018, 769, 521–531. [Google Scholar] [CrossRef]
  25. Marzec, A.; Radeck, M.; Maziarzc, W.; Kusior, A.; Pędzich, Z. Structural, optical and electrical properties of nanocrystalline TiO2, SnO2 and their composites obtained by the sol–gel method. J. Eur. Ceram. Soc. 2016, 36, 2981–2989. [Google Scholar] [CrossRef]
  26. Lyson-Sypien, B.; Czapla, A.; Lubecka, M.; Kusior, E.; Zakrzewska, K.; Radecka, M.; Kusior, A.; Balogh, A.G.; Lauterbach, S.; Kleebe, H.-J. Gas sensing properties of TiO2–SnO2 nanomaterials. Sens. Actuators B 2013, 187, 445–454. [Google Scholar] [CrossRef]
  27. Kusior, A.; Radecka, M.; Zych, Ł.; Zakrzewska, K.; Reszka, A.; Kowalski, B.J. Sensitization of TiO2/SnO2 nanocomposites for gas detection. Sens. Actuators B 2013, 189, 251–259. [Google Scholar] [CrossRef]
  28. Zeng, W.; Liu, T.; Zang, Z. Sensitivity improvement of TiO2-doped SnO2 to volatile organic compounds. Phys. E 2010, 43, 633–638. [Google Scholar] [CrossRef]
  29. Ahn, K.; Pham-Cong, D.; Choi, H.S.; Jeong, S.-Y.; Cho, J.H.; Kim, J.; Kim, J.-P.; Bae, J.-S.; Cho, C.-R. Bandgap-designed TiO2/SnO2 hollow hierarchical nanofibers: Synthesis, properties, and their photocatalytic mechanism. Curr. Appl. Phys. 2016, 16, 251–260. [Google Scholar] [CrossRef]
  30. Chizhov, A.S.; Rumyantseva, M.N.; Vasiliev, R.B.; Filatova, D.G.; Drozdov, K.A.; Krylov, I.V.; Marchevsky, A.V.; Karakulina, O.M.; Abakumov, A.M.; Gaskov, A.M. Visible light activation of room temperature NO2 gas sensors based on ZnO, SnO2 and In2O3 sensitized with CdSe quantum dots. Thin Solid Films 2016, 618, 253–262. [Google Scholar] [CrossRef]
  31. Rumyantseva, M.N.; Gaskov, A.M.; Rosman, N.; Pagnier, T.; Morante, J.R. Raman surface vibration modes in nanocrystalline SnO2 prepared by wet chemical methods: Correlations with the gas sensors performances. Chem. Mater. 2005, 17, 893–901. [Google Scholar] [CrossRef]
  32. Güntner, A.T.; Pineau, N.J.; Chie, D.; Krumeich, F.; Pratsinis, S.E. Selective sensing of isoprene by Ti-doped ZnO for breath diagnostics. J. Mater. Chem. B 2016, 4, 5358–5366. [Google Scholar] [CrossRef]
  33. Rumyantseva, M.N.; Gaskov, A.M. Chemical modification of nanocrystalline metal oxides: Effect of the real structure and surface chemistry on the sensor properties. Russ. Chem. Bull. 2008, 57, 1106–1125. [Google Scholar] [CrossRef]
  34. Vorobyeva, N.A.; Rumyantseva, M.N.; Vasiliev, R.B.; Kozlovskiy, V.F.; Soshnikova, Y.M.; Filatova, D.G.; Zaytsev, V.B.; Zaytseva, A.V.; Gaskov, A.M. Doping effects on electrical and optical properties of spin-coated ZnO thin films. Vacuum 2015, 114, 198–204. [Google Scholar] [CrossRef]
  35. Korotcenkov, G.; Cho, B.K. Metal oxide composites in conductometric gas sensors: Achievements and challenges. Sens. Actuators B 2017, 244, 182–210. [Google Scholar] [CrossRef]
  36. Miodownik, M.; Holm, E.A.; Hassold, G.N. Highly parallel computer simulation of particle pinning: Zener vindicated. Scripta Mater. 2000, 42, 1173–1177. [Google Scholar] [CrossRef]
  37. Socrates, G. Infrared and Raman Characteristic Group Frequencies: Tables and Charts, 3rd ed.; John Wiley & Sons Ltd.: West Sussex, UK, 2001. [Google Scholar]
  38. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds. Part A: Theory and Applications in Inorganic Chemistry, 6th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2009. [Google Scholar]
  39. Kumar, A.; Rout, L.; Dhaka, R.S.; Samala, S.L.; Dash, P. Design of a graphene oxide-SnO2 nanocomposite with superior catalytic efficiency for the synthesis of β-enaminones and β-enaminoesters. RSC Adv. 2015, 5, 39193–39204. [Google Scholar] [CrossRef]
  40. Tang, Y.; Fu, S.; Zhao, K.; Xie, G.; Teng, L. Synthesis of TiO2 nanofibers with adjustable anatase/rutile ratio from Ti sol and rutile nanoparticles for the degradation of pollutants in wastewater. Ceram. Int. 2015, 41, 13285–13293. [Google Scholar] [CrossRef]
  41. Karthick, S.N.; Hemalatha, K.V.; Seo, H.; Ludeman, D.; Kim, J.-K.; Prabakar, K.; Kim, H.-J. Titanium oxide prepared by polymer gel assisted combustion method for dye-sensitized solar cell. Curr. Appl. Phys. 2011, 11, 127–130. [Google Scholar] [CrossRef]
  42. León, A.; Reuquen, P.; Garín, C.; Segura, R.; Vargas, P.; Zapata, P.; Orihuela, P.A. FTIR and Raman characterization of TiO2 nanoparticles coated with polyethylene glycol as carrier for 2-methoxyestradiol. Appl. Sci. 2017, 7, 49. [Google Scholar] [CrossRef]
  43. Peercy, P.S.; Morosin, B. Pressure and temperature dependences of the Raman-active phonons in SnO2. Phys. Rev. B 1973, 7, 2779–2786. [Google Scholar] [CrossRef]
  44. Garcia-Tecedor, M.; Maestre, D.; Cremades, A.; Piqueras, J. Growth and characterization of Cr doped SnO2 microtubes with resonant cavity modes. J. Mater. Chem. C 2016, 4, 5709–5716. [Google Scholar] [CrossRef]
  45. Caoshui, X.; Yonghong, X.; Hong, Z.; Yuheng, Z.; Yulong, L. Investigation of Raman spectrum for nano-SnO2. Sci. China (Ser. A). 1997, 40, 1222–1227. [Google Scholar]
  46. Ferreira, C.S.; Santos, P.L.; Bonacin, J.A.; Passos, R.R.; Pocrifka, L.A. Rice husk reuse in the preparation of SnO2/SiO2 Nanocomposite. Mat. Res. 2015, 18, 639–643. [Google Scholar] [CrossRef]
  47. Zuo, J.; Xu, C.; Liu, X.; Wang, C.; Wang, C.; Hu, Y.; Qian, Y. Study of the Raman spectrum of nanometer SnO2. J. Appl. Phys. 1994, 75, 1835–1836. [Google Scholar] [CrossRef]
  48. Abello, L.; Bochu, B.; Gaskov, A.; Koudryavtseva, S.; Lucazeau, G.; Roumyantseva, M. Structural characterization of nanocrystalline SnO2 by X-ray and Raman spectroscopy. J. Solid State Chem. 1998, 135, 78–85. [Google Scholar] [CrossRef]
  49. Dieguez, A.; Romano-Rodriguez, A.; Vila, A.; Morante, J.R. The complete Raman spectrum of nanometric SnO2 particles. J. Appl. Physics. 2001, 90, 1550–1557. [Google Scholar] [CrossRef]
  50. Liu, L.Z.; Li, T.H.; Wu, X.L.; Shen, J.C.; Chu, P.K. Identification of oxygen vacancy types from Raman spectra of SnO2 nanocrystals. J. Raman Spectrosc. 2012, 43, 1423–1426. [Google Scholar] [CrossRef]
  51. Loudon, R. The Raman effect in crystals. Adv. Phys. 1964, 13, 423–482. [Google Scholar] [CrossRef]
  52. Verma, R.; Mantri, B.; Ramphal, A.K.S. Shape control synthesis, characterizations, mechanisms and optical properties of large scaled metal oxide nanostructures of ZnO and TiO2. Adv. Mater. Lett. 2015, 6, 324–333. [Google Scholar] [CrossRef]
  53. Gupta, S.K.; Desai, R.; Jha, P.K.; Sahoo, S.P.; Kirin, D. Titanium dioxide synthesized using titanium chloride: Size effect study using Raman and Photoluminescence. J. Raman Spectrosc. 2010, 41, 350–355. [Google Scholar] [CrossRef]
  54. Bersani, D.; Lottici, P.P.; Ding, X.Z. Phonon confinement effects in the Raman scattering by TiO2 nanocrystals. Appl. Phys. Lett. 1998, 72, 73–75. [Google Scholar] [CrossRef]
  55. Zhang, W.F.; Zhang, M.S.; Yin, Z.; Chen, Q. Photoluminescence in anatase titanium dioxide nanocrystals. Appl. Phys. B 2000, 70, 261–265. [Google Scholar] [CrossRef]
  56. Tompsett, G.A.; Bowmaker, G.A.; Cooney, R.P.; Metson, J.B.; Rodgers, K.A.; Seakins, J.M. The Raman spectrum of brookite, TiO2 (Pbca, Z = 8). J. Raman Spectrosc. 1995, 26, 57–62. [Google Scholar] [CrossRef]
  57. Rezaee, M.; Khoie, S.M.M.; Liu, K.H. The role of brookite in mechanical activation of anatase-to-rutile transformation of nanocrystalline TiO2: An XRD and Raman spectroscopy investigation. CrystEngComm 2011, 13, 5055–5061. [Google Scholar] [CrossRef]
  58. Nikodemski, S.; Dameron, A.A.; Perkins, J.D.; O’Hayre, R.P.; Ginley, D.S.; Berry, J.J. The role of nanoscale seed layers on the enhanced performance of niobium doped TiO2 thin films on glass. Sci. Rep. 2016, 6, 32830. [Google Scholar] [CrossRef]
  59. Zhang, H.; Banfield, J.F. Understanding polymorphic phase transformation behavior during growth of nanocrystalline aggregates: Insights from TiO2. J. Phys. Chem. B 2000, 104, 3481–3487. [Google Scholar] [CrossRef]
  60. Zhu, K.-R.; Zhang, M.-S.; Hong, J.-M.; Yin, Z. Size effect on phase transition sequence of TiO2 nanocrystal. Mater. Sci. Eng. A 2005, 403, 87–93. [Google Scholar] [CrossRef]
  61. Fan, S.W.; Srivastava, A.K.; Dravid, V.P. UV-activated room-temperature gas sensing mechanism of polycrystalline ZnO. Appl. Phys. Lett. 2009, 95, 142106-1–142106-3. [Google Scholar] [CrossRef]
  62. Xu, Y.; Schoonen, M.A.A. The absolute energy positions of conduction and valence bands of selected semiconducting minerals. Am. Miner. 2000, 85, 543–556. [Google Scholar] [CrossRef]
  63. Oguchiand, T.; Fujishima, A. Photocatalytic degradation of gaseous formaldehyde using TiO2 film. Environ. Sci. Technol. 1998, 32, 3831–3833. [Google Scholar]
  64. Sun, S.; Ding, J.J.; Bao, J.; Gao, C.; Qi, Z.; Li, C. Photocatalytic oxidation of gaseous formaldehyde on TiO2: An in situ DRIFTS study. Catal. Lett. 2010, 137, 239–246. [Google Scholar] [CrossRef]
  65. Varechkina, E.N.; Rumyantseva, M.N.; Vasiliev, R.B.; Konstantinova, E.A.; Gaskov, A.M. UV-VIS photoconductivity of nanocrystalline tin oxide. J. Nanoelectron. Op. 2012, 7, 623–628. [Google Scholar] [CrossRef]
  66. Rumyantseva, M.; Nasriddinov, A.; Vladimirova, S.; Fedorova, O.; Tokarev, S.; Krylov, I.; Drozdov, K.; Baranchikov, A.; Gaskov, A. Photosensitive organic-inorganic hybrid materials for room temperature gas sensor applications. Nanomaterials 2018, 8, 671. [Google Scholar] [CrossRef]
  67. Popova, G.Y.; Budneva, A.A.; Andrushkevich, T.V. Identification of adsorption forms by IR spectroscopy for formaldehyde and formic acid on K3PMo12O40. React. Kinet. Catal. Lett. 1997, 61, 353–362. [Google Scholar] [CrossRef]
  68. Busca, G.; Lamotte, J.; Lavalley, J.-C.; Lorenzelli, V. FT-IR study of the adsorption and transformation of formaldehyde on oxide surfaces. J. Am. Chem. Soc. 1987, 109, 5197–5202. [Google Scholar] [CrossRef]
  69. Gomes, J.R.B.; Gomes, J.A.N.F. A theoretical study of dioxymethylene, proposed as intermediate in the oxidation of formaldehyde to formate over copper. Surf. Sci. 2000, 446, 283–293. [Google Scholar] [CrossRef]
  70. Chen, D.; Qu, Z.; Sun, Y.; Gao, K.; Wang, Y. Identification of reaction intermediates and mechanism responsible for highly active HCHO oxidation on Ag/MCM-41 catalysts. Appl. Catal. B 2013, 142–143, 838–848. [Google Scholar] [CrossRef]
  71. Xu, B.Y.; Shang, J.; Zhu, T.; Tang, X.Y. Heterogeneous reaction of formaldehyde on the surface of γ-Al2O3. Atmos. Environ. 2011, 45, 3569–3575. [Google Scholar] [CrossRef]
  72. Gao, H.W.; Yan, T.X.; Zhang, C.B.; He, H. Theoretical experimental analysis on vibrational spectra of formate species adsorbed on Cu-Al2O3 catalyst. J. Mol. Str. THEOCHEM 2008, 857, 38–43. [Google Scholar] [CrossRef]
  73. Zhang, C.; He, H.; Tanaka, K. Catalytic performance and mechanism of a Pt/TiO2 catalyst for the oxidation of formaldehyde at room temperature. Appl. Catal. B 2006, 65, 37–43. [Google Scholar] [CrossRef]
  74. Chen, B.; Shi, C.; Crocker, M.; Wang, Y.; Zhu, A. Catalytic removal of formaldehyde at room temperature over supported gold catalysts. Appl. Catal. B 2013, 132–133, 245–255. [Google Scholar] [CrossRef]
  75. Guo, J.; Lin, C.; Jiang, C.; Zhang, P. Review on noble metal-based catalysts for formaldehyde oxidation at room temperature. Appl. Surf. Sci. 2019, 475, 237–255. [Google Scholar] [CrossRef]
  76. Huang, K.; Kong, L.; Yuan, F.; Xie, C. In situ diffuse reflectance infrared Fourier transform spectroscopy study of formaldehyde adsorption and reactions on nano γ-Fe2O3 films. Appl. Surf. Sci. 2013, 270, 405–410. [Google Scholar] [CrossRef]
  77. Lin, S.D.; Cheng, H.K.; Hsiao, T.C. In situ DRIFTS study on the methanol oxidation by lattice oxygen over Cu/ZnO catalyst. J. Mol. Catal. A 2011, 342–343, 35–40. [Google Scholar] [CrossRef]
  78. Rumyantseva, M.N.; Makeeva, E.A.; Badalyan, S.M.; Zhukova, A.A.; Gaskov, A.M. Nanocrystalline SnO2 and In2O3 as materials for gas sensors: The relationship between microstructure and oxygen chemisorption. Thin Solid Films 2009, 518, 1283–1288. [Google Scholar] [CrossRef]
  79. Tharsika, T.; Haseeb, A.S.M.A.; Akbar, S.A.; Sabri, M.; Hoong, W.Y. Enhanced ethanol gas sensing properties of SnO2-core/ZnO-shell nanostructures. Sensors 2014, 14, 14586–14600. [Google Scholar] [CrossRef]
  80. Dobrokhotov, V.; Larin, A. Multisensory gas chromatography for field analysis of complex gaseous mixtures. ChemEngineering 2019, 3, 13. [Google Scholar] [CrossRef]
  81. Miller, D.R.; Akbar, S.A.; Morris, P.A. Nanoscale metal oxide-based heterojunctions for gas sensing: A review. Sens. Actuators B 2014, 204, 250–272. [Google Scholar] [CrossRef]
  82. Buckeridge, J.; Butler, K.T.; Catlow, C.R.A.; Logsdail, A.J.; Scanlon, D.O.; Shevlin, S.A.; Woodley, S.M.; Sokol, A.A.; Walsh, A. Polymorph Engineering of TiO2: Demonstrating How Absolute Reference Potentials Are Determined by Local Coordination. Chem. Mater. 2015, 27, 3844–3851. [Google Scholar] [CrossRef]
  83. Li, X.S.; Li, W.Z.; Wang, H.L. Enhancement of hydrogen spillover by surface labile oxygen species on oxidized Pt/TiO2 catalyst. Catal. Lett. 1995, 32, 31–42. [Google Scholar]
  84. Wang, J.A.; Cuan, A.; Salmones, J.; Nava, N.; Castillo, S.; Morán-Pineda, M.; Rojas, F. Studies of sol–gel TiO2 and Pt/TiO2 catalysts for NO reduction by CO in an oxygen-rich condition. Appl. Surf. Sci. 2004, 230, 94–105. [Google Scholar] [CrossRef]
  85. Jiang, X.; Ding, G.; Lou, L.; Chen, Y.; Zheng, X. Catalytic activities of CuO/TiO2 and CuO-ZrO2/TiO2 in NO + CO reaction. J. Molec. Catal. A 2004, 218, 187–195. [Google Scholar]
  86. Kemmler, J.A.; Pokhrel, S.; Birkenstock, J.; Schowalter, M.; Rosenauer, A.; Bârsan, N.; Weimar, U.; Mädler, L. Quenched, nanocrystalline In4Sn3O12 high temperature phase for gas sensing applications. Sens. Actuators B 2012, 161, 740–747. [Google Scholar] [CrossRef]
  87. Park, H.J.; Choi, N.J.; Kang, H.; Jung, M.Y.; Park, J.W.; Park, K.H.; Lee, D.S. A ppb-level formaldehyde gas sensor based on CuO nanocubes prepared using a polyol process. Sens. Actuators B 2014, 203, 282–288. [Google Scholar] [CrossRef]
  88. Fang, F.; Bai, L.; Sun, H.; Kuang, Y.; Sun, X.; Shi, T.; Song, D.; Guo, P.; Yang, H.; Zhang, Z.; et al. Hierarchically porous indium oxide nanolamellas with ten-parts-per-billion-level formaldehyde-sensing performance. Sens. Actuators B 2015, 206, 714–720. [Google Scholar] [CrossRef]
  89. Lv, P.; Tang, Z.A.; Yu, J.; Zhang, F.T.; Wei, G.F.; Huang, Z.X.; Hu, Y. Study on a micro-gas sensor with SnO2–NiO sensitive film for indoor formaldehyde detection. Sens. Actuators B. 2008, 132, 74–80. [Google Scholar] [CrossRef]
  90. Zhang, Y.; Liu, Q.; Zhang, J.; Zhu, Q.; Zhu, Z. A highly sensitive and selective formaldehyde gas sensor using a molecular imprinting technique based on Ag–LaFeO3. J. Mater. Chem. C 2014, 2, 10067–10072. [Google Scholar] [CrossRef]
  91. Zhang, Y.M.; Lin, J.Y.T.; Chen, L.; Zhang, J.; Zhu, Z.Q.; Liu, Q.J. A high sensitivity gas sensor for formaldehyde based on silver doped lanthanum ferrite. Sens. Actuators B 2014, 190, 171–176. [Google Scholar] [CrossRef]
  92. Chung, F.-C.; Zhu, Z.; Luo, P.-Y.; Wu, R.-J.; Li, W. Au@ZnO core–shell structure for gaseous formaldehyde sensing at room temperature. Sens. Actuators B 2014, 199, 314–319. [Google Scholar] [CrossRef]
  93. Chen, T.; Liu, Q.J.; Zhou, Z.L.; Wang, Y.D. The fabrication and gas-sensing characteristics of the formaldehyde gas sensors with high sensitivity. Sens. Actuators B 2008, 131, 301–305. [Google Scholar] [CrossRef]
  94. Wang, Y.Q.H.; Chen, H.; Lin, Z.; Dai, K. Highly selective n-butanol gas sensor based on mesoporous SnO2 prepared with hydrothermal treatment. Sens. Actuators B 2014, 201, 153–159. [Google Scholar] [CrossRef]
  95. Hu, R.; Wang, J.; Chen, P.; Hao, Y.; Zhang, C.; Li, X. Preparation of Cd-loaded In2O3 hollow nanofibers by electro spinning and improvement of formaldehyde sensing performance. J. Nanomater. 2014, 2014, 7. [Google Scholar] [CrossRef]
  96. Liu, D.; Pan, J.; Tang, J.; Liu, W.; Bai, S.; Luo, R. Ag decorated SnO2 nanoparticles to enhance formaldehyde sensing properties. J. Phys. Chem. Solids 2019, 124, 36–43. [Google Scholar] [CrossRef]
  97. Du, H.; Wang, J.; Su, M.; Yao, P.; Zheng, Y.; Yu, N. Formaldehyde gas sensor based on SnO2/In2O3 hetero-nanofibers by a modified double jets electro spinning process. Sens. Actuators B 2012, 166, 746–752. [Google Scholar] [CrossRef]
  98. Castro-Hurtado, I.; Herrán, J.; Mandayo, G.G.; Castaño, E. SnO2-nanowires grown by catalytic oxidation of tin sputtered thin films for formaldehyde detection. Thin Solid Films 2012, 520, 4792–4796. [Google Scholar] [CrossRef]
  99. Sun, P.; Zhou, X.; Wang, C.; Shimanoe, K.; Lu, G.; Yamazoe, N. Hollow SnO2/α-Fe2O3 spheres with a double-shell structure for gas sensors. J. Mater. Chem. A 2014, 2, 1302–1308. [Google Scholar] [CrossRef]
  100. Wang, J.; Zhang, P.; Qi, J.-Q.; Yao, P.-J. Silicon-based micro-gas sensors for detecting formaldehyde. Sens. Actuators B 2009, 136, 399–404. [Google Scholar] [CrossRef]
  101. Lee, C.-Y.; Chiang, C.-M.; Wang, Y.-H.; Ma, R.-H. A self-heating gas sensor with integrated NiO thin-film for formaldehyde detection. Sens. Actuators B 2007, 122, 503–510. [Google Scholar] [CrossRef]
  102. Zhang, G.; Zhang, S.; Yang, L.; Zou, Z.; Zeng, D.; Xie, C. La2O3-sensitized SnO2 nanocrystalline porous film gas sensors and sensing mechanism toward formaldehyde. Sens. Actuators B 2013, 188, 137–146. [Google Scholar] [CrossRef]
  103. Castro-Hurtado, I.; Herrán, J.; Mandayo, G.G.; Castaño, E. Studies of influence of structural properties and thickness of NiO thin films on formaldehyde detection. Thin Solid Films 2011, 520, 947–952. [Google Scholar] [CrossRef]
  104. Peng, L.; Zhai, J.; Wang, D.; Zhang, Y.; Wang, P.; Zhao, Q.; Xie, T. Size- and photoelectric characteristics-dependent formaldehyde sensitivity of ZnO irradiated with UV light. Sens. Actuators B 2010, 148, 66–73. [Google Scholar] [CrossRef]
  105. Yao, P.J.; Wang, J.; Chu, W.L.; Hao, Y.W. Preparation and characterization of La1-xSrxFeO3 materials and their formaldehyde gas-sensing properties. J. Mater. Sci. 2013, 48, 441–450. [Google Scholar] [CrossRef]
  106. Wang, J.; Liu, L.; Cong, S.-Y.; Qi, J.-Q.; Xu, B.-K. An enrichment method to detect low concentration formaldehyde. Sens. Actuators B 2008, 134, 1010–1015. [Google Scholar] [CrossRef]
  107. Chung, F.-C.; Wu, R.-J.; Cheng, F.-C. Fabrication of a Au@SnO2 core–shell structure for gaseous formaldehyde sensing at room temperature. Sens. Actuators B 2014, 190, 1–7. [Google Scholar] [CrossRef]
  108. Li, Y.; Chen, N.; Deng, D.; Xing, X.; Xiao, X.; Wang, Y. Formaldehyde detection: SnO2 microspheres for formaldehyde gas sensor with high sensitivity, fast response/recovery and good selectivity. Sens. Actuators B 2017, 238, 264–273. [Google Scholar] [CrossRef]
  109. Zeng, W.; Liu, T.; Wang, Z.; Tsukimoto, S.; Saito, M.; Ikuhara, Y. Selective detection of formaldehyde gas using a Cd-doped TiO2-SnO2 sensor. Sensors 2009, 9, 9029–9038. [Google Scholar] [CrossRef]
  110. Tian, H.; Fan, H.; Li, M.; Ma, L. Zeolitic Imidazolate Framework Coated ZnO Nanorods as Molecular Sieving to Improve Selectivity of Formaldehyde Gas Sensor. ACS Sens. 2016, 1, 243–250. [Google Scholar] [CrossRef]
  111. Li, Y.; Jin, H.; Sun, G.; Zhang, B.; Luo, N.; Lin, L.; Bala, H.; Cao, J.; Zhang, Z.; Wang, Y. Synthesis of novel porous ZnO octahedrons and their improved UV-light activated formaldehyde-sensing performance by Au decoration. Phys. E 2019, 106, 40–44. [Google Scholar] [CrossRef]
  112. Mondal, B.; Mukherjee, K.; Das, P. Facile synthesis of pseudo-peanut shaped hematite iron oxide nano-particles and their promising ethanol and formaldehyde sensing characteristics. RSC Adv. 2014, 4, 31879–31886. [Google Scholar]
  113. Han, N.; Wu, X.; Zhang, D.; Shen, G.; Liu, H.; Chen, Y. CdO activated Sn-doped ZnO for highly sensitive, selective and stable formaldehyde sensor. Sens. Actuators B 2011, 152, 324–329. [Google Scholar] [CrossRef]
  114. Xie, C.; Xiao, L.; Hu, M.; Bai, Z.; Xia, X.; Zeng, D. Fabrication and formaldehyde gas-sensing property of ZnO–MnO2 coplanar gas sensor arrays. Sens. Actuators B 2010, 145, 457–463. [Google Scholar] [CrossRef]
  115. Han, N.; Tian, Y.; Wu, X.; Chen, Y. Improving humidity selectivity in formaldehyde gas sensing by a two-sensor array made of Ga-doped ZnO. Sens. Actuators B 2009, 138, 228–235. [Google Scholar] [CrossRef]
  116. Zhang, H.; Song, P.; Han, D.; Wang, Q. Synthesis and formaldehyde sensing performance of LaFeO3 hollow nanospheres. Phys. E 2014, 63, 21–26. [Google Scholar] [CrossRef]
  117. Chu, X.; Chen, T.; Zhang, W.; Zheng, B.; Shui, H. Investigation on formaldehyde gas sensor with ZnO thick film prepared through microwave heating method. Sens. Actuators B 2009, 142, 49–54. [Google Scholar] [CrossRef]
  118. Güntner, A.T.; Abegg, S.; Wegner, K.; Pratsinis, S.E. Zeolite membranes for highly selective formaldehyde sensors. Sens. Actuators B 2018, 257, 916–923. [Google Scholar] [CrossRef]
  119. Gulevich, D.; Rumyantseva, M.; Gerasimov, E.; Marikutsa, A.; Krivetskiy, V.; Shatalova, T.; Khmelevsky, N.; Gaskov, A. Nanocomposites SnO2/SiO2 for CO gas sensors: Microstructure and reactivity in the interaction with the gas phase. Materials 2019, 12, 1096. [Google Scholar] [CrossRef]
  120. Gulevich, D.G.; Marikutsa, A.V.; Rumyantseva, M.N.; Fabrichnyi, P.B.; Shatalova, T.B.; Gaskov, A.M. Detection of carbon monoxide in humid air with double-layer structures based on semiconducting metal oxides and silicalite. Russ. J. Appl. Chem. 2018, 91, 1671–1679. [Google Scholar] [CrossRef]
  121. Chen, E.-X.; Yang, H.; Zhang, J. Zeolitic Imidazolate Framework as Formaldehyde Gas Sensor. Inorg. Chem. 2014, 53, 5411–5413. [Google Scholar] [CrossRef]
  122. Krivetskiy, V.; Efitorov, A.; Arkhipenko, A.; Vladimirova, S.; Rumyantseva, M.; Dolenko, S.; Gaskov, A. Selective detection of individual gases and CO/H2 mixture at low concentrations in air by single semiconductor metal oxide sensors working in dynamic temperature mode. Sens. Actuators B 2018, 254, 502–513. [Google Scholar] [CrossRef]
Figure 1. Micro-hotplate scheme: (a) front side; (b) back side; (c) Current–voltage (I-V) characteristics of the sensors.
Figure 1. Micro-hotplate scheme: (a) front side; (b) back side; (c) Current–voltage (I-V) characteristics of the sensors.
Sensors 19 03182 g001
Figure 2. Schematic illustration of sensor measurements setup. (1) DC power supply QJ2002C (Ningbo JiuYuan Electronic, Zhejiang, China); (2) Time relay REV-114 (Novatek-Electro, Moscow, Russia); (3) Teflon airtight and light-tight chamber (laboratory-made); (4) UV LED (SMD-5050 XML, Epileds Co. LTD, Taiwan, China); (5) Control PC; (6) Analyte gas bottle (4.6 ppm of HCOH in N2, certified gas mixture, Monitoring, Russia); (7) Pure air generator («Granat-Engineering» Co.Ltd, Moscow, Russia); (8) Mass Flow Controller EL-FLOW F-221M (Bronkhorst, Ruurlo, The Netherlands); (9) Humidifier Cellkraft P-2 (Cellkraft AB, Stockholm, Sweden); (10) electronic module providing control of sensor heating and high-precision measurement of the resistance of the sensitive layer (MGA-2-1, Senseria, Tomsk, Russia).
Figure 2. Schematic illustration of sensor measurements setup. (1) DC power supply QJ2002C (Ningbo JiuYuan Electronic, Zhejiang, China); (2) Time relay REV-114 (Novatek-Electro, Moscow, Russia); (3) Teflon airtight and light-tight chamber (laboratory-made); (4) UV LED (SMD-5050 XML, Epileds Co. LTD, Taiwan, China); (5) Control PC; (6) Analyte gas bottle (4.6 ppm of HCOH in N2, certified gas mixture, Monitoring, Russia); (7) Pure air generator («Granat-Engineering» Co.Ltd, Moscow, Russia); (8) Mass Flow Controller EL-FLOW F-221M (Bronkhorst, Ruurlo, The Netherlands); (9) Humidifier Cellkraft P-2 (Cellkraft AB, Stockholm, Sweden); (10) electronic module providing control of sensor heating and high-precision measurement of the resistance of the sensitive layer (MGA-2-1, Senseria, Tomsk, Russia).
Sensors 19 03182 g002
Figure 3. (a) TEM and (b) Lattice resolved TEM image of SnO2 powder used as a substrate in ALD synthesis; (c) Lattice-resolved TEM image of SnTi-1 nanocomposite.
Figure 3. (a) TEM and (b) Lattice resolved TEM image of SnO2 powder used as a substrate in ALD synthesis; (c) Lattice-resolved TEM image of SnTi-1 nanocomposite.
Sensors 19 03182 g003
Figure 4. XRD patterns of nanocrystalline SnO2 and TiO2@SnO2 composites and references for SnO2 cassiterite (ICDD 41-1445), TiO2 anatase (ICDD 21-1272) and TiO2 brookite (ICDD 29-1360).
Figure 4. XRD patterns of nanocrystalline SnO2 and TiO2@SnO2 composites and references for SnO2 cassiterite (ICDD 41-1445), TiO2 anatase (ICDD 21-1272) and TiO2 brookite (ICDD 29-1360).
Sensors 19 03182 g004
Figure 5. (a) FT-IR spectra of nanocrystalline SnO2 and TiO2@SnO2 composites normalized to the intensity of (O–Sn–O) oscillations; (b) Normalized Raman spectra of nanocrystalline SnO2 and TiO2@SnO2 composites.
Figure 5. (a) FT-IR spectra of nanocrystalline SnO2 and TiO2@SnO2 composites normalized to the intensity of (O–Sn–O) oscillations; (b) Normalized Raman spectra of nanocrystalline SnO2 and TiO2@SnO2 composites.
Sensors 19 03182 g005
Figure 6. Resistance of the TiO2@SnO2 nanocomposites in the temperature range 300–50 °C under periodic change of the gas phase composition (a) in dark conditions; (b) under periodic UV (λmax = 365 nm) illumination.
Figure 6. Resistance of the TiO2@SnO2 nanocomposites in the temperature range 300–50 °C under periodic change of the gas phase composition (a) in dark conditions; (b) under periodic UV (λmax = 365 nm) illumination.
Sensors 19 03182 g006
Figure 7. (a) Temperature dependencies of sensor response of the TiO2@SnO2 nanocomposites toward 0.6 ppm HCHO in air in dark conditions (left) and under UV periodic illumination (right); (b) Temperature dependencies of effective photoresponse SPh of the TiO2@SnO2 nanocomposites in background air.
Figure 7. (a) Temperature dependencies of sensor response of the TiO2@SnO2 nanocomposites toward 0.6 ppm HCHO in air in dark conditions (left) and under UV periodic illumination (right); (b) Temperature dependencies of effective photoresponse SPh of the TiO2@SnO2 nanocomposites in background air.
Sensors 19 03182 g007
Figure 8. Temperature dependencies of response time τ 90 (left) and recovery time τ 90 * (right) during detection of 0.6 ppm HCHO in air in dark conditions.
Figure 8. Temperature dependencies of response time τ 90 (left) and recovery time τ 90 * (right) during detection of 0.6 ppm HCHO in air in dark conditions.
Sensors 19 03182 g008
Figure 9. In situ DRIFT spectra of the unmodified SnO2 (SnTi-0) (a) and SnTi-2 nanocomposite (b) during HCHO adsorption at room temperature.
Figure 9. In situ DRIFT spectra of the unmodified SnO2 (SnTi-0) (a) and SnTi-2 nanocomposite (b) during HCHO adsorption at room temperature.
Sensors 19 03182 g009
Figure 10. In situ DRIFT spectra of the SnTi-2 nanocomposite obtained during the heating in dry air at T = 300 °C after formaldehyde adsorption at room temperature.
Figure 10. In situ DRIFT spectra of the SnTi-2 nanocomposite obtained during the heating in dry air at T = 300 °C after formaldehyde adsorption at room temperature.
Sensors 19 03182 g010
Figure 11. (a) Estimated band alignment of the SnO2 (cassiterite) and TiO2 (anatase) phases (left) and SnO2/TiO2 n-n heterocontact (right); (b) TPR-H2 profiles of unmodified SnO2 (SnTi-0) and TiO2@SnO2 nanocomposites.
Figure 11. (a) Estimated band alignment of the SnO2 (cassiterite) and TiO2 (anatase) phases (left) and SnO2/TiO2 n-n heterocontact (right); (b) TPR-H2 profiles of unmodified SnO2 (SnTi-0) and TiO2@SnO2 nanocomposites.
Sensors 19 03182 g011
Figure 12. Resistance of the TiO2@SnO2 nanocomposites at T = 150 °C under stepwise change of the HCHO concentration (a) in dark conditions; (b) under periodic UV (λmax = 365 nm) illumination. The numbers in the figure correspond to HCHO concentration (ppm).
Figure 12. Resistance of the TiO2@SnO2 nanocomposites at T = 150 °C under stepwise change of the HCHO concentration (a) in dark conditions; (b) under periodic UV (λmax = 365 nm) illumination. The numbers in the figure correspond to HCHO concentration (ppm).
Sensors 19 03182 g012
Figure 13. (a) Calibration curves for the HCHO detection using TiO2@SnO2 nanocomposites at T = 150 °C under stepwise change of the HCHO concentration in dark conditions (left) and under UV periodic illumination (right). (b) Comparison of the values of the sensor response of SnTi-2 nanocomposite (triangles) with the literature data [8,9,10,11,20,21,22,23,28,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117] (squares), Ref. [7] (diamonds) and Ref. [86] (circles). The color of the symbol corresponds to the measurement temperature. For correct comparison all the data were recalculated as S = (Rair/Rgas - 1)*100%.
Figure 13. (a) Calibration curves for the HCHO detection using TiO2@SnO2 nanocomposites at T = 150 °C under stepwise change of the HCHO concentration in dark conditions (left) and under UV periodic illumination (right). (b) Comparison of the values of the sensor response of SnTi-2 nanocomposite (triangles) with the literature data [8,9,10,11,20,21,22,23,28,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117] (squares), Ref. [7] (diamonds) and Ref. [86] (circles). The color of the symbol corresponds to the measurement temperature. For correct comparison all the data were recalculated as S = (Rair/Rgas - 1)*100%.
Sensors 19 03182 g013
Figure 14. Cross sensitivity of SnO2 (SnTi-0, black), SnTi-1 (red), SnTi-2 (blue) and SnTi-3 (green) nanocomposites toward HCHO (circles), benzene (squares) and acetone (triangles) at 150 °C.
Figure 14. Cross sensitivity of SnO2 (SnTi-0, black), SnTi-1 (red), SnTi-2 (blue) and SnTi-3 (green) nanocomposites toward HCHO (circles), benzene (squares) and acetone (triangles) at 150 °C.
Sensors 19 03182 g014
Table 1. Elemental and phase composition of SnO2 reference sample and TiO2@SnO2 nanocomposites with crystallite size dXRD of detected phases.
Table 1. Elemental and phase composition of SnO2 reference sample and TiO2@SnO2 nanocomposites with crystallite size dXRD of detected phases.
SampleNumber of ALD Cycles[Ti]/([Ti]+[Sn]), mol.%Phase Composition
PhasesdXRD, nm
SnTi-000SnO2 cassiterite9 ± 1
SnTi-150020 ± 2SnO2 cassiterite5.1 ± 0.5
TiO2 brookite13 ± 1
TiO2 anatasen/d
SnTi-2100024 ± 2SnO2 cassiterite4.5 ± 0.5
TiO2 brookite16 ± 2
TiO2 anatasen/d
SnTi-3200039 ± 3SnO2 cassiterite3.5 ± 0.5
TiO2 anatase22 ± 2
Table 2. Assignment of Raman vibrational modes (cm−1).
Table 2. Assignment of Raman vibrational modes (cm−1).
SnTi-0SnTi-1SnTi-2SnTi-3
137 (B1g, cassiterite)145 (Eg, anatase)145 (Eg, anatase)145 (Eg, anatase)
309 (IR Eu, cassiterite)198 (Eg, anatase)198 (Eg, anatase)198 (Eg, anatase)
352 (IR Eu, cassiterite)255 (A1g, brookite)255 (A1g, brookite)
444 (IR B1u, cassiterite)285 (B3g, brookite)285 (B3g, brookite)
481 (Eg, cassiterite)397 (B1g, anatase)397 (B1g, anatase)397 (B1g, anatase)
563 (surface mode)443 (B3g, brookite)443 (B3g, brookite)
631 (A1g, cassiterite)516 (A1g+B1g, anatase)516 (A1g+B1g, anatase)516 (A1g+B1g, anatase)
773 (B2g, cassiterite)581 (B2g, brookite)581 (B2g, brookite)
635 (Eg, anatase)635 (Eg, anatase)635 (Eg, anatase)
Table 3. Assignment of IR bands (cm−1) for the intermediates of HCHO oxidation.
Table 3. Assignment of IR bands (cm−1) for the intermediates of HCHO oxidation.
SampleFormaldehyde HCHODOM H2COO-Formate HCOO-Other Bands
SnTi-01202 τ (CH2)1050 ν(CO)2838 ν(CH)2337 CO2
1742 ν(CO)1107 ρ(CH2)2890 ν(CH)2365 CO2
1766 ν(CO)1145 ρ(CH2)
30102936 δ (CH2)
SnTi-21768 ν(CO)1066 ν(CO)1344 νs(COO)2340 CO2
1107 ρ(CH2)1346 νs(COO)3550 ν(OH)
1146 ρ(CH2)1382 δ(CH)3630 ν(OH)
1165 ρ(CH2)1553 νass(COO)
1450 δs(CH2)1592 νass(COO)
1478 δs (CH2)2590
2930 δ (CH2)2868 ν(CH)
Table 4. Sensor characteristics of unmodified SnO2 and TiO2@SnO2 nanocomposites in HCHO detection at 150 °C.
Table 4. Sensor characteristics of unmodified SnO2 and TiO2@SnO2 nanocomposites in HCHO detection at 150 °C.
SampleS (to 0.06 ppm)LDL, ppbRav, Ohmτ90, sτ*90, s
DarkUVDarkUVDarkUV
SnTi-01.341.179301.2·1051.1·105360780
SnTi-11.491.209321.2·1067.4·105120660
SnTi-21.721.2615326.5·1072.2·107120660
SnTi-31.481.2110305.8·1063.5·106120680
Table 5. Literature data on sensor response S = (Rair/Rgas - 1)*100% of metal oxide semiconductor gas sensor in formaldehyde detection.
Table 5. Literature data on sensor response S = (Rair/Rgas - 1)*100% of metal oxide semiconductor gas sensor in formaldehyde detection.
MaterialHCHO conc., ppmResponse ST, oCRef.
Si-SnO2 FSP a films0.0036.5400[7]
In4Sn3O12 thin films0.02110350[86]
CuO nanocubes0.054300[87]
SnO2-Au0.052300[8]
In2O3 nanolamellas0.084300[88]
TiO2 UV activation0.167RT[22]
SnO2-NiO0.318300[89]
Ag-LaFeO30.5240040[90]
Ag-LaFeO31240090[91]
TiO2 hollow microspheres53900RT[23]
Au@ZnO core-shell structure5957RT[92]
CdO-In2O310990095[93]
SnO2 mesoporous109900150[94]
TiO2 macroporous106900RT[20]
Cd-In2O3 hollow fibers101500280[95]
Ag-SnO2 nanoparticles101340125[96]
SnO2/In2O3 nanofibers10650375[97]
SnO2 nanowires10145270[98]
SnO2/α-Fe2O3 hollow spheres10100250[99]
Pd-SnO2 thin films1055250[100]
NiO thin films1050300[101]
TiO2 nanotube arrays107RT[21]
La2O3-SnO2 thin films18.75400250[102]
NiO thin films20330340[103]
ZnO UV activated5010700RT[104]
TiO2-SnO2503000360[28]
La1-xSrxFeO3502500200[105]
SnO2-MWCNT b50300250[106]
Au@SnO2 core-shell structure50190RT[107]
SnO2 microspheres1003730200[108]
Cd-TiO2/SnO21001400320[109]
ZnO/ZIF c1001100300[110]
Au-ZnO octahedrons100660RT[111]
γ-Fe2O3100500320[77]
α-Fe2O310052325[112]
CdO/Sn-ZnO200200000200[113]
ZnO-MnO22002600320[114]
Ga-ZnO20513400[115]
LaFeO3 hollow nanospheres5001900260[116]
ZnO100099600210[117]
a Flame spray pyrolysis; b Multiwall carbon nanotubes; c Zeolitic Imidazolate Framework.

Share and Cite

MDPI and ACS Style

Nasriddinov, A.; Rumyantseva, M.; Marikutsa, A.; Gaskov, A.; Lee, J.-H.; Kim, J.-H.; Kim, J.-Y.; Kim, S.S.; Kim, H.W. Sub-ppm Formaldehyde Detection by n-n TiO2@SnO2 Nanocomposites. Sensors 2019, 19, 3182. https://0-doi-org.brum.beds.ac.uk/10.3390/s19143182

AMA Style

Nasriddinov A, Rumyantseva M, Marikutsa A, Gaskov A, Lee J-H, Kim J-H, Kim J-Y, Kim SS, Kim HW. Sub-ppm Formaldehyde Detection by n-n TiO2@SnO2 Nanocomposites. Sensors. 2019; 19(14):3182. https://0-doi-org.brum.beds.ac.uk/10.3390/s19143182

Chicago/Turabian Style

Nasriddinov, Abulkosim, Marina Rumyantseva, Artem Marikutsa, Alexander Gaskov, Jae-Hyoung Lee, Jae-Hun Kim, Jin-Young Kim, Sang Sub Kim, and Hyoun Woo Kim. 2019. "Sub-ppm Formaldehyde Detection by n-n TiO2@SnO2 Nanocomposites" Sensors 19, no. 14: 3182. https://0-doi-org.brum.beds.ac.uk/10.3390/s19143182

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop