DOI:
10.1039/D5RA05533D
(Paper)
RSC Adv., 2025,
15, 38750-38761
SnO2 quantum dot decoration of CuO nanoparticles with enhanced NO2 and H2 gas sensing response via p–n heterojunction interfaces
Received
30th July 2025
, Accepted 29th September 2025
First published on 15th October 2025
Abstract
Design and fabrication of heterostructures has emerged as a powerful strategy to improve gas sensing performances compared to single materials counterparts. In this work, we report an innovative CuO-based nanostructure decorated with SnO2 quantum dots (QDs) for the detection of NO2 and H2 gases. Here, CuO serves as the base material while SnO2 QDs are used as the decorating phase: an inversion of the conventional architecture where SnO2 is typically the host and CuO the modifier. The composite exhibits higher sensitivity compared to pristine CuO and SnO2, showing state-of-the-art performances in terms of relative responses (RRs) in the 20 ppb to 1 ppm range and 10 ppm to 250 ppm for NO2 and H2 respectively, with excellent stability and reproducibility. Moreover, the SnO2-QDs/CuO operates at a low working temperature (i.e. 100 °C), offering significant advantages in terms of energy efficiency and material stability. The observed enhancements are attributed to the optimized heterointerface, increased active surface area, and modulation of the charge carrier induced by the p–n heterojunctions. These results highlight the potential of reverse-configured SnO2/CuO as a versatile platform for improved, low-temperature gas sensors with high sensitivity.
1. Introduction
The detection of toxic and explosive gases such as nitrogen oxide (NO2) and hydrogen (H2) is a critical requirement in the fields of health and environmental monitoring and industrial safety. Specifically, NO2 is an air pollutant primarily originating from combustion processes associated with vehicle traffic and industrial activities, which can cause serious respiratory diseases even at low concentrations.1,2 On the other hand, H2 gas, being colourless and odourless, poses a significant risk due to its wide flammability range (4–75% in air), low ignition energy (0.017 mJ) and high combustion heat (142 kJ per g H2), negatively influencing its exploitation as a promising fuel for the transition to a low carbon economy.3,4 To address these challenges, the development of sensitive and reliable gas sensors for detection of trace levels of NO2 and H2 remains an active area of research. Thin films of metal oxide semiconductors (MOX) have been widely employed as chemo-resistive gas sensors due to their high sensitivity, low cost and simple fabrication processes. Typically, n-type MOX materials such as SnO2, ZnO and In2O3,5–7 and p-type CuO, NiO and Cr2O3 counterparts8–10 have been widely studied for gas sensing applications. However, one major limitation is their high operating temperature, often exceeding 200 °C, which increases power consumption and hinders their integration in portable or wearable devices.11–13
Recent studies have demonstrated that carefully engineered oxide heterostructures can significantly enhance charge separation and gas adsorption processes.14,15 In particular, a promising strategy is represented by the design of heterojunctions combining p-type and n-type metal oxides, leveraging the formation of p–n junctions to improve gas sensing at lower temperatures.16,17 Among p-type materials, CuO stands out due to its excellent surface reactivity, stability and ability to form hierarchical nanostructures in the form of flakes, nanorods and nanofibers.8,18–20 On the other hand, SnO2, being one of the first MOX sensor ever studied,21 is well known for its strong interactions with both oxidizing and reducing gases.22,23 The combination of CuO and SnO2 has been explored extensively, typically in the configuration where SnO2 serves as the base material and CuO is introduced as the secondary phase.24–27 These structures, showing typical n-type response, have been extensively utilized for H2S sensing on account of the reversible conversion of CuO in CuS, which significantly reduces the resistance.28 In this work, we propose a reverse composite configuration, employing CuO as the primary sensing matrix and decorating it with SnO2 quantum dots (QDs). This architecture leverages the predominant p-type response of CuO-based interfaces to oxidizing gases like NO2,29 while promoting intimate contact between the SnO2 QDs edges and the underlying CuO surface, resulting in enhanced interfacial charge transfer and gas sensing performance.30 Remarkably, this SnO2-QDs/CuO architecture exhibits superior sensing performance towards both NO2 and H2 at a reduced operating temperature of 100 °C. Specifically, needle-like CuO was synthesized via microwave irradiation of Cu2(OH)2CO3, while SnO2 QDs were produced using a colloidal solution method. The formation of nanoscale p–n junctions at the SnO2/CuO interfaces, enhancing charge carrier separation and modulation of the depletion layer, is responsible to improve gas sensing response.17 Furthermore, the choice of SnO2 in quantum dots form introduces significant advantages. Due to its nanoscale size and high surface-to-volume ratio, SnO2 provides more active sites for gas adsorption, easing faster charge transfer kinetics.31,32 Moreover, the quantum confinement effect also modulates the electronic properties of the heterojunction, enhancing sensitivity at low temperatures (i.e. 100 °C). Overall, this study provides a novel perspective on CuO-based heterojunctions, proposing a simple and scalable synthesis route that make this approach suitable for practical application, demonstrating enhanced NO2 and H2 sensing performances at low temperatures, thus offering a promising alternative to conventional single-phase or inversely configured systems.
2. Experimental
2.1 Synthesis of CuO nanostructures
The hydrothermal synthesis of Cu2(OH)2CO3 was performed according to a reported procedure:33 4.2 ml of 0.4 M copper(II) acetate and 4.2 ml of 0.8 M urea were dissolved into 3.3 ml DI water. After 30 min stirring, the mixture was transferred into a 40 ml Teflon-lined stainless-steel autoclave and maintained at 120 °C for 4 h. After that, the autoclave was naturally cooled down to room temperature and the obtained product was washed 5 times with DI water and 3 times with absolute ethanol and dried at 50 °C for 1 h. Then 0.1 g of Cu2(OH)2CO3 were added to 20 ml of DI water, and the obtained suspension was irradiated with microwave at 560 W for 5 minutes. After collecting and washing, a brown precipitate was obtained. Finally, the powder was annealed in a muffle furnace at 400 °C for 2 h with a heating rate of 1 °C min−1 to obtain crystalline CuO. A detailed flowchart of the process is reported in SI S1a.
2.2 Synthesis of SnO2 QDs
50 ml 0.025 M ammonia solution 30% was added dropwise into 50 ml 0.05 M SnCl4·5H2O solution and kept stirring at 80 °C in oil bath for 3 h.34 At the end of the reaction, the resulting white gel-like product was collected by centrifugation at 5000 rpm and washed five times with DI water to remove any residual precursors. The precipitate was then dried overnight at 60 °C. For further characterization, the dried SnO2 product was redispersed in ethanol for deposition and analysis. A detailed flowchart of the process is reported in SI S1b.
2.3 Decorating CuO nanoparticles with SnO2-QDs
The optimized procedure for SnO2 decoration of CuO flakes was established through systematic variation of precursor concentration and order of addition (see SI, Fig. S2). In the final protocol, 25 mg of synthesized crystalline CuO were added to 20 ml of DI water; then, 2.6 mg SnCl4·5H2O powder and 0.3 ml of 0.05 M ammonia solution were successively incorporated to the dispersion. After mixing, obtained dispersion was heated at 80 °C using an oil bath and maintained under stirring for 3 h. The obtained product was washed three times by centrifugation at 5000 rpm, followed by a final centrifugation at the same speed to exchange the solvent with ethanol, facilitating solvent evaporation after deposition and enhancing both microstructural and electrical characterization. A detailed flowchart of the process is reported in SI S1c.
Notably, the hydrothermal synthesis of Cu2(OH)2CO3 typically afforded a yield of ∼43% (∼130 mg of Cu2(OH)2CO3 from ∼300 mg of copper(II) acetate). In contrast, both the SnO2 and the SnO2/CuO syntheses showed essentially quantitative conversion of the Sn precursor, as no secondary phases or unreacted material were detected by our structural analyses.
2.4 Material characterization
The crystal structure was analysed through the X-ray diffraction Grazing Incidence X-ray Diffraction (GI-XRD) by XRD-PANAnalytical X’PERT Pro using Cu Kα1 radiation (λ = 1.5406 Å) with an incident angle of 0.8°. FTIR spectra were obtained by Thermo Nicolet Nexus 870, operating in the 400–4000 cm−1 spectral range. Thermogravimetric and differential thermal analysis (TG-DTA) was performed in air atmosphere using a Linseis L81-I. The morphology was studied by Transmission Electron Microscopy Philips CM100 operating at 100 kV and Scanning Electron Microscopy Gemini SEM working at 5 kV. HRTEM was performed using a Talos F200S. The optical absorption and reflectance spectra were measured using a PerkinElmer LAMBDA 1050+ UV-vis-near-infrared (NIR) spectrophotometer and in diffuse reflectance spectroscopy (DRS) configuration respectively, and the Kubelka–Munk function was utilized.
2.5 Electrical characterization
CuO, SnO2 and SnO2-QDs/CuO based films were deposited by spin coating on Si/Si3N4 substrates, with comb-like Pt interdigitated electrodes (30 μm apart) and back side heaters and inserted in a 500 cm3 Teflon chamber for chemoresistive gas sensing characterization. By stepwise tuning the current in the back side heater, sensor's substrates are almost instantaneously kept and maintained at the selected operating temperature. Electrical responses were collected using an automated volt–amperometric system (Agilent 34970A), measuring electrical resistance of the film at operating temperatures from 25 to 150 °C and different environments. Gas concentrations were obtained by diluting NO2 (10 ppm in air) and H2 (500 ppm in air) certified mixtures (Nippon gases-IT) with synthetic dry air utilizing mass flow controllers (MKS147), setting the total flow rate at 500 sccm per min. 50% relative humidity (RH%) was obtained by mixing dry with water-saturated air at 25 °C and measuring RH% at 25 °C before injection into the test chamber (Thermohygrometer – Hannah Instruments).
To characterize and compare gas responses properties of the different samples, the following definitions apply: (i) baseline resistance (BLR): the resistance in dry air at equilibrium; (ii) relative response (RR): defined as (RAir/RGas) or (RGas/RAir) depending on the oxidizing/reducing nature of the gases; (iii) sensor's sensitivity (S): the slope of the calibration curve (i.e. RR vs. gas concentration); (iv) response time (τADS) and recovery time (τDES) defined as the time required for the resistance to reach 90% of the equilibrium value after injecting the gas and the time needed to return to 10% above the original value in air, respectively.
3. Results and discussion
We have applied three different chemical syntheses to prepare: (i) needle-like CuO nanoparticles; (ii) SnO2-QDs; and (iii) SnO2-QDs decorated needle-like CuO (SnO2-QDs/CuO). The detailed flowcharts of processes (i)–(iii) are reported in SI S1. Notably, while the CuO and SnO2 syntheses followed established literature routes,33,34 the synthesis of SnO2-QDs/CuO was the result of an optimization process involving both the relative amount of precursors and the order of their addition, as extensively described in SI S2. All material interfaces, deposited on dedicated Si3N4 substrates provided with Pt finger-type electrodes and a backside heater, have been investigated as NO2 and H2 sensors under dry and humid air background conditions.
3.1 Synthesis and microstructural characterization
3.1.1 Needle-like CuO nanoparticles. Following the procedure described in the Experimental section and Fig. 1a and SI S1a, the hydrothermal synthesis utilizing copper(II) acetate and urea solution leads to the formation of round-shaped agglomerated-particles with chemical composition Cu2(OH)2CO3. Specifically, Scanning Electron Microscopy (SEM) (Fig. 1b), exhibits the formation of hierarchical microspheres made of “flower-like” assembled-nanosheets, growing radially from the core of the particle, with similar features as those described in literature.33 According to Fig. 1c, the chemical composition of the synthesized “flower-like” structures (black line), correspond to Cu2(OH)2CO3 as observed by comparing the Fourier Transform Infrared (FT-IR) spectrum of the synthesized structures with that of a commercial malachite powder (orange line). X-ray Diffraction (XRD) analysis (Fig. 1d) confirms that Cu2(OH)2CO3 flower-like particles are highly crystalline (black line of Fig. 1d), consistent with JCPDS Card No. 00-041-1390 corresponding to malachite.
 |
| | Fig. 1 (a) Schematics of the synthesis process: after hydrothermal synthesis of Cu2(OH)2CO3 at 120 °C for 4 h, powder was dispersed in DI water and irradiated with microwave (MW) at 560 W for 5 min and annealed at 400 °C for 2 h to obtain crystalline needle-like CuO; (b) SEM image of Cu2(OH)2CO3 after hydrothermal synthesis; (c) comparison of the FT-IR spectra of: commercial malachite (Cu2(OH)2CO3, orange line), Cu2(OH)2CO3 after hydrothermal synthesis (black line), microwave irradiated (pink line) and annealed (green line) Cu2(OH)2CO3; (d) XRD spectra of Cu2(OH)2CO3, microwave irradiated and annealed Cu2(OH)2CO3 with associated JCPDS cards; (e) TEM image of needle-like Cu2(OH)2CO3 after MW; (f) thermogravimetric (TG) and differential thermal analysis (DTA) plots of MW Cu2(OH)2CO3 powder heated in air at 5 °C min−1 from 25 to 800 °C. Black and red lines refer to TG and DTA signals, respectively; (g) TEM image of needle-like crystalline CuO after annealing. | |
The “flower-like” Cu2(OH)2CO3 agglomerated structure (Fig. 1a), after microwave (MW) treatment at 560 W for 5 minutes, separates into “needle-like” free nanoparticles as those displayed in Fig. 1e, while maintaining the same chemical composition of Cu2(OH)2CO3 (pink line of Fig. 1c). Surprisingly, the XRD pattern of the microwave synthesized needle-like Cu2(OH)2CO3 particles, reveals the absence of distinct diffraction peaks (pink line of Fig. 1d), suggesting that the microwave treatment induces amorphization of the particles, while preserving their original Cu2(OH)2CO3 chemical composition. Finally, the amorphous “needle-like” Cu2(OH)2CO3 particles were subjected to annealing in dry air, to promote the formation of crystalline CuO. Preliminary simultaneous thermogravimetric (TG) and differential thermal analysis (DTA) technique was utilized to determine the most favourable annealing temperature.
Specifically, heating the amorphous “needle-like” Cu2(OH)2CO3 at 5 °C min−1 from 25 to 400 °C in a simultaneous TG-DTA apparatus, the thermogravimetric (TG) curve (black curve of Fig. 1f) yields a weight loss of −28.5% with a maximum decomposition rate at 305 °C (red line of the DTA signal of Fig. 1f). Remarkably, the measured weight loss of −28.5%, satisfactorily matches, within the experimental error, the theoretical weight loss of −30.5%, corresponding to the complete conversion of Cu2(OH)2CO3 into CuO, based on the following reaction:
| Cu2(OH)2CO3(s) → 2CuO(s) + CO2(g) + H2O(g) |
Consequently, after oven annealing at 400 °C for 2 h in static air, the “needle-like” amorphous Cu2(OH)2CO3 particles are isomorphically converted into “needle-like” CuO nanoparticles (Fig. 1g). Moreover, XRD characterization (green line of Fig. 1d) exhibits that CuO nanoparticles are highly crystalline (JCPDS Card 00-041-0254) and almost pure, as confirmed by FTIR analysis (green line of Fig. 1c). In conclusion, congruent with previous research,33,35 we synthesized pure and highly crystalline needle-like CuO nanoparticles to be utilized as scaffolds for SnO2-QDs decoration.
3.1.2 SnO2-QD synthesis and SnO2-QD/CuO nanoparticles decoration. The SnO2 Quantum Dots (SnO2-QD) synthesis,34 shown in Fig. 2a and described in Experimental section and SI S1b, yields agglomerated spherical SnO2 nanoparticles (TEM Fig. 2c) with average radii of R = 1.9 ± 0.1 nm and diameter-size distribution shown in the inset of Fig. 2c. XRD analysis of the synthesized SnO2-QDs nanoparticles, exhibits the formation of low-crystalline SnO2 structures, as demonstrated by the presence of broad diffraction peaks in the red curve of Fig. 2d, matching rutile phase of tetragonal tin oxide (JCPDS Card No. 00-001-0657). According to this procedure, we successfully synthesized SnO2 nanoparticles, with average radii (R = 1.9 ± 0.1 nm) smaller than Bohr's exciton radius (2.7 nm) in SnO2-QDs,34,36 attesting the capability of the synthesized SnO2-QDs to yield quantum confinements effects.
 |
| | Fig. 2 (a) Synthesis of SnO2 QDs and (b) SnO2-QDs/CuO; (c) TEM image of SnO2 QDs with inset reporting particle size distribution; (d) XRD spectra of SnO2 QDs and SnO2-QDs/CuO with associated JCPDS cards; (e) TEM image of SnO2-QDs/CuO. | |
The SnO2-QDs decoration of the CuO needle-like nanoparticles to yield SnO2-QD/CuO, was carried out according to an optimized procedure (Fig. 2b and SI S1, S2) by mixing previously prepared needle-like CuO in 20 ml DI water with 2.6 mg SnCl4·5H2O and 0.3 ml of 0.05 M ammonia solution (Fig. 2b) to yield, after centrifugation and washing, decorated SnO2-QDs/CuO. TEM analysis of the SnO2-QDs/CuO, shown in Fig. 2e, exhibits the formation of small spherical SnO2 QDs nanoparticles of ≈4 nm size over the edge of the CuO needle-like structures. The XRD pattern of the SnO2-QD/CuO (blue line of Fig. 2d) confirms the presence of crystalline CuO phase, but it does not display any discernible reflections belonging to SnO2-QDs, possibly on account of a limited quantity of the SnO2-QDs phase as respect to the CuO matrix. HRTEM characterizations of the CuO needle-like, SnO2-QDs and SnO2-QDs/CuO, shown in Fig. 3, highlight that CuO needle-like nanoparticles are highly crystalline (Fig. 3a–d), comprising well-ordered lattice fringes with interplanar spacing of 0.25 nm (Fig. 3d), corresponding to the (002) plane of monoclinic CuO.33
 |
| | Fig. 3 (a)–(d) TEM characterization of not decorated needle-like CuO: (a) low resolution TEM showing CuO microstructure, (b)–(d) high resolution TEM showing well-ordered crystalline structure (e)–(h) TEM characterization of as synthesized SnO2 QDs: (e) low resolution TEM showing dots microstructure, (f)–(h) high resolution TEM showing the small crystalline domains of SnO2 QDs; (i) high resolution TEM of SnO2-QDs/CuO and EDX maps (j) and (k) with indicated the elemental composition. | |
In a similar way, HRTEM of SnO2 QDs (Fig. 3e–h), display the occurrence of crystalline domains (i.e. inside the white box of Fig. 3f), with average diameter's size smaller than 5 nm and interplanar spacing of 0.33 nm (Fig. 3h), consistent with the (110) plane of tetragonal SnO2. The effectiveness of the SnO2 QDs decoration of CuO needle-like nanoparticles is finally confirmed by HRTEM of Fig. 3i and by the associated EDX elemental map of the SnO2-QDs/CuO composite (Fig. 3j and k), attesting a congruent distribution of Cu (cyan), Sn (red) and O (green) elements over the CuO needle-like scaffold.
Finally in Fig. 4 we have reported a schematization of the SnO2-QDs/CuO model structure. Congruently with HRTEM characterization (see Fig. 3), SnO2-QDs (white) are discretely distributed as isolated spots over the CuO underlying surface (black). The whole structure comprises a network morphology of needle like CuO nanoparticles decorated with SnO2-QDs.
 |
| | Fig. 4 TEM image with superimposed a schematic illustration of the SnO2-QDs/CuO model structure. The SnO2 quantum dots (white) are evenly dispersed as isolated spots on the CuO surface (black). The overall structure consists of a network-like arrangement of needle-shaped CuO nanoparticles, which are decorated with SnO2 quantum dots. | |
3.2 Optical properties characterization
The optical absorbance of CuO needle-like nanoparticles, SnO2-QDs and SnO2-QDs/CuO was measured by UV-vis-near-infrared spectrophotometer as shown in Fig. 5a. Associated bandgaps, reported in Fig. 5b, were determined by DRS applying the Kubelka–Munk method.37 The CuO needle-like nanoparticles absorption curve (green line of Fig. 5a), exhibits a maximum located at ≈380 nm, attributed to electrons band-gap excitation in CuO.38 SnO2-QDs (red line), on the other hand, highlight an absorption edge at ≈280 nm, blue shifted compared to that of bulk SnO2 (345 nm).39 According to literature,40 the blue shift may represent the quantum confinement effect associated to the synthesized SnO2-QDs. Finally, the absorption of SnO2-QDs/CuO represented by the blue line in Fig. 5a, does not show any change in the wavelength position compared to that of CuO, but its absorbance is increased considerably. This behavior can be associated to the change in CuO band gap due to SnO2 decoration and to the formation of surface defects.25
 |
| | Fig. 5 Optical characterization of CuO, SnO2 QDs and SnO2-QDs/CuO. (a) UV-vis absorption spectra (0.5 mg ml−1 in ethanol) and (b) Tauc plot with highlighted the measured bandgaps determination, respectively. | |
The bandgaps of CuO needle-like nanoparticles, SnO2-QDs and SnO2-QDs/CuO displayed in Fig. 5b were extrapolated by Tauc’s plot41 using Kubelka–Munk37 function from diffuse reflectance spectra (E1):
| | |
(F(R)hv)1/γ = B(hv − Eg)
| (E1) |
here
h [kg m
2 s
−1] is Planck's constant,
ν [s
−1] is the frequency of the incident electromagnetic radiation,
Eg [eV] is the optical bandgap energy,
B is a constant and
γ = 1/2, corresponding to a direct band transition for CuO, SnO
2-QDs and SnO
2-QDs/CuO. The measured bandgap (BG) of 1.8 eV found for CuO needle-like nanoparticles (green line of
Fig. 5b), is different from that of bulk CuO (1.24 eV) and closer to the reported BG value of 1.73 eV of CuO nanopetals.
42 The 4.1 eV measured bandgap of SnO
2-QDs (red line of
Fig. 5b), which is consistently higher than that of bulk SnO
2 (3.6 eV),
34 can be possibly explained considering the QDs nature of the synthesized SnO
2-QDs nanoparticles as respect to bulk SnO
2. To better clarify the nature of this mismatch, we estimated the SnO
2-QDs bandgap using Brus's variational method,
43 according to
eqn (E2),
| |
 | (E2) |
where
Eg is the bulk band gap energy (3.6 eV),
ℏ is reduced Planck's constant,
μ is the effective reduced mass expressed as

and
μ may be replaced by the electron effective mass (

), since

(

and

are the electron and hole effective masses, respectively).
44 R is the radius of QDs,
e is a charge of an electron (1.602 × 10
−19 C) and
ε is the dielectric constant (for SnO
2,
ε = 14). Hence, the calculated band gap energy using
eqn (E2) is ∼4.0 eV, which is close to the measured band gap of 4.1 eV (
Fig. 5b), confirming the quantum confinement effect occurring in the synthesized SnO
2 QDs.
34 Finally, the bandgap of SnO
2-QDs/CuO is determined as 2.0 eV, which is higher than that of CuO needle-like nanoparticles (
Eg = 1.8 eV), with this increment again related to the presence of SnO
2-QDs,
24 as previously discussed.
3.3 Gas sensing characterization
The semiconducting nature of the CuO needle-like nanoparticles, SnO2-QDs and SnO2-QDs/CuO heterojunction is attested by the decrease/increase of the baseline resistance (i.e. BLR – the resistance at equilibrium in dry air), when increasing/decreasing the operating temperature (OT) between 25 °C and 150 °C, as shown in Fig. 6a. CuO nanoparticles and SnO2-QDs/CuO (green and blue lines) show higher baseline resistances compared to pristine SnO2-QDs (red line). The BLR of CuO nanoparticles is displayed departing from 75 °C, since at T < 75 °C the material's resistance exceeds the instrumental capabilities. Oppositely, the base line resistance (BLR) of SnO2-QDs is smaller than the others with BLR's modulation limited in the range 4–20 kΩ. The lower BLR of the SnO2-QDs film can be related to a size quantization effect, which has recently been demonstrated to be responsible for the increase in conductivity, when the size of the SnO2 nanoparticles is lowered within the nanometer scale.45 The SnO2-QDs/CuO shows BLR values that are in between those of CuO nanoparticles and SnO2-QDs, demonstrating the effectiveness of the SnO2 decoration to tune the film conductivity.
 |
| | Fig. 6 (a) BLR modulation of CuO (green), SnO2 (red) and SnO2-QDs/CuO (blue) measured in dry air under increasing/decreasing the OT in the 25–150–25 °C range; (b) gas sensing response to 1 ppm NO2 and (c) to 100 ppm H2 at different OTs (25–150 °C); (d) and (e) comparison of the normalized gas response to 1 ppm NO2 and 100 ppm H2 at 100 °C OT; (f) SnO2-QDs/CuO reproducibility response to NO2 (1 ppm) and H2 (100 ppm) at 100 °C OT; (g) and (h) dynamic electrical responses of the SnO2-QDs/CuO in dry air at an OT of 100 °C to NO2 (20 ppb to 1 ppm) and H2 (10 ppm to 250 ppm); (i) SnO2-QDs/CuO log/log calibration plots at 100 °C OT of the sensor's signal (i.e., RR = Ra/Rg or Rg/Ra) vs. NO2 and H2 gas concentrations. | |
To evaluate the gas sensing properties, concentration ranges of 0.020–1 ppm for NO2 and 10–250 ppm for H2, were selected. This choice reflects the significantly higher electron affinity of NO2 (2.3 eV) compared to H2 (0.18 eV),46,47 which leads to a more pronounced sensor signal variation for NO2 at equivalent concentrations.
To evaluate the optimal operating temperature (OT) of each sensor, films were exposed to low concentrations of NO2 (1 ppm) and H2 (100 ppm) in dry air by varying the temperature in the range of 25–150 °C, as shown in Fig. 6b and c. At 25 °C OT, the base line resistance moves beyond 109 Ω exceeding the instrumental capabilities. By stepwise increasing the operating temperature from 25 °C to 150 °C (Fig. 6b and c) the base line resistance of all sensors decreases, consistent with their semiconducting nature (see also Fig. 6a). Specifically, as shown in Fig. 6b, starting from 75 °C the introduction of 1 ppm NO2 into the test chamber (indicated by the grey rectangle at the bottom), causes a noticeable decrease of the resistance for both the CuO (green) and SnO2-QDs/CuO (blue) sensors. Opposedly, base line resistance of SnO2 remains largely unaffected upon exposure to NO2. Conversely, as indicated in Fig. 6c, when 100 ppm H2 is introduced into the chamber, the resistance of both CuO (green) and SnO2-QDs/CuO (blue) increases. It turns out that both CuO nanoparticles and SnO2-QDs/CuO exhibit p-type behavior, associated to a decrease/increase of the resistance upon NO2/H2 exposure, congruently with previous studies on CuO48 and CuO-based heterostructure sensors.49,50 Opposedly, SnO2 quantum dots (QDs) demonstrate typical n-type response, characteristic of SnO2 metal oxide sensors,51 with a negligible resistance increase/decrease to NO2/H2.
Considering now that the onset gas sensing temperature to NO2/H2 of the SnO2-QDs/CuO heterojunction is 50/100 °C, while that of SnO2-QD and CuO is 100 °C (Fig. 6b and c), we set the operating temperature (OT) at 100 °C, as a balance between the sensor's signal amplitude (here referred as the relative response, RR) and a fast and reversible baseline recovery. Specifically, sensors' electrical responses at 100 °C to 1 ppm NO2 and 100 ppm H2 in dry air are shown in Fig. 6d and e. We found that at 1 ppm NO2, the RRs values (Ra/Rg or Rg/Ra, depending on the p/n nature of the sensor) are 1.6, 2.5, and 1.2 for CuO, SnO2-QDs/CuO and SnO2-QDs (with an estimated standard deviation of ±0.1 calculated over a set of 5 identical measurements). RRs to 100 ppm H2 are slightly smaller: 1.2, 1.7, and 1.1, confirming the superior performances of the SnO2-QDs/CuO to detect both NO2 and H2. As a concluding remark, the SnO2-QD/CuO shows excellent reverse capability to measure NO2 and H2 with fast and reversible base line recovery, as displayed in Fig. 6f.
Fig. 6g and h show the dynamic resistance changes of the SnO2-QDs/CuO to NO2 in the range 20 ppb to 1 ppm and H2 (10–250 ppm) in dry air background at 100 °C OT. SnO2-QDs/CuO shows a remarkable modulation of the sensor's signal to increasing NO2 and H2 gas concentrations, with excellent recovery of the BLR following each step of gas/dry air purge. The log–log calibration plots of the sensor's signal (i.e., RR = Ra/Rg or Rg/Ra) vs. NO2 and H2 gas concentrations shown in Fig. 6i, yield gas sensitivities (S), as represented by the slope of the calibration lines, SNO2 = (0.21 ± 0.01) [ppm]−1 and SH2 = (0.06 ± 0.01) [ppm]−1. The limits of detections (LOD) for NO2 and H2 were determined by extrapolating the calibration lines in Fig. 6i to the point where the response ratio (RR) equals 1, yielding LOD(NO2) = 12 ppb and LOD(H2) = 15 ppb respectively. The better NO2 dynamic responses (RRs) and sensitivities (S) with respect to H2, confirm the stronger tendency of NO2 molecules to interact with the SnO2-QD/CuO surface as respect to less electronegative H2, as it will be discussed in the next paragraph.
Retrieving from literature data, Table 1 compares normalized relative responses – to 1 ppm NO2 and 100 ppm H2 (when available) of recently published CuO-based composites for gas sensing. Apart from Pd–CuO/rGO interfaces operating at 25 °C, our SnO2-QD/CuO yields comparable performances, eventually obtained at lower operating temperature as respect to the others.
Table 1 Comparison of the NO2 and H2 gas sensing response of CuO based composites normalized at 1 ppm NO2 and 100 ppm H2 (where applicable). Relative Response (RRs) indicated as: Rg/Ra if Rg > Ra; Ra/Rg if Rg < Ra
| Sensing material |
Gas concentration |
Relative response RR [—] |
OT [°C] |
Ref. |
| NO2gas |
| SnOx/CuO |
1 ppm |
2.5 |
250 |
49 |
| CuO/SnOx |
1 ppm |
1.5 |
250 |
49 |
| CuO/ZnO |
1 ppm |
1.05 |
250 |
20 |
| CuO/SnO2 nanoflowers |
1 ppm |
10 |
100 |
29 |
| Pd–CuO/rGO |
1 ppm |
4.5 |
25 |
52 |
| SnO2-QDs/CuO |
1 ppm |
2.5 |
100 |
This work |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
| H2gas |
| CuO/Fe2O4 |
500 ppm |
1.8 |
400 |
53 |
| Nb2O5/CuO |
0.5% |
2 |
300 |
54 |
| NiO/CuO |
20 ppm |
8 |
150 |
55 |
| In2O3/CuO |
400 ppm |
1.3 |
350 |
56 |
| TiO2/CuO/Cu2O |
100 ppm |
6 |
350 |
57 |
| SnO2-QDs/CuO |
100 ppm |
1.7 |
100 |
This work |
In addition, SnO2-QDs/CuO sensor's adsorption/desorption times (τads/τdes) to H2 are faster with respect to NO2. Comparing τads/τdes to NO2 (Fig. 7a) and H2 (Fig. 7d), NO2 adsorption/desorption times are approximately two-fold those of H2, with desorption times always bigger than adsorption for both gases. Cross sensitivity tests were carried out to investigate the effect of humidity as interfering gas to the NO2 and H2 response, as shown in Fig. 7b and e. The standardized cross-sensitivity procedure comprises three steps. In a first step (panel i) SnO2-QD/CuO is exposed to 1 ppm NO2 and 100 ppm H2 in dry air; in a second step (panel ii) dry air background is replaced by 50% Relative Humidity (RH) humid-air while exposing the sensor to 1 ppm NO2 and 100 ppm H2, finally a third step (panel iii) is carried out in the same conditions of (panel i) to check for short term reproducibility. According to Fig. 7b and e, BLR of the SnO2-QD/CuO rapidly increases as soon as 50% RH is introduced, in line with p-type sensors response to humidity at operating at temperature higher than 25 °C.58–60 Significantly, relative response to 1 ppm NO2 increases from 2.5 ± 0.1 in dry air, to 2.8 ± 0.1 in 50% RH. That to 100 ppm H2 decreases from 1.7 ± 0.1 to 1.4 ± 0.1 respectively. This behavior is consistent with previous findings about a possible synergistic/antisynergistic effect of water vapor in the presence of oxidizing/reducing gases.23,51 Selectivity response (SR) shown in SI Fig. S3, highlights the superior ability of SnO2-QDs/CuO sensor to detect NO2 and H2 compared to 100 ppm NH3, 1% ethanol and 1% acetone. Notably, SR differs from humidity cross response (Fig. 7b and e), since it compares the sensor's ability to preferentially respond to a specific gas (i.e. 1 ppm NO2 target gas) while ignoring other gases or organic vapours under similar experimental conditions.
 |
| | Fig. 7 (a) and (d) Adsorption and desorption times of the SnO2-QD/CuO in dry air at an OT of 100 °C to NO2 (20 ppb to 1 ppm) and H2 (10 ppm to 250 ppm); (b) NO2 and (e) H2 cross-sensitivity to 50% relative humidity (RH). Each panel of figures (b)–(e) comprises: first panel (i), the response to 1 ppm NO2 (100 ppm H2) in dry air, second panel (ii), response 1 ppm NO2 (100 ppm H2) in 50% RH background, third panel (iii), response to 1 ppm NO2 (100 ppm H2), as to first panel, to check repeatability; (c) and (f) reproducibility and baseline recovery measured by exposing the SnO2-QDs/CuO to both pulse and cumulative NO2 (20 ppb to 1 ppm) and H2 (10–250 ppm) concentrations. | |
Finally, the short term reproducibility of the electrical response to NO2 and H2 to both pulsed (on/off) and cumulative (increasing/decreasing) concentration modes is shown in Fig. 7c and f. Specifically, under pulsed conditions, the baseline resistance (BLR) overall recovers its initial value after each desorption cycle in dry air, with a slight displacement from the BLR (as highlighted by Δads for both gases), corresponding to higher gases' concentration. In a similar way, under cumulative stepwise adsorption/desorption modes, the NO2/H2 gases resistances taken at 1 ppm/100 ppm, perfectly matches the corresponding ones in pulse mode.
4. Gas sensing mechanism
The SnO2-QDs/CuO exhibits improved sensing performances with respect to its single counterparts, showing a prevailing a p-type response. Apart from the formation of p–n junctions, indeed SnO2 quantum dots significantly enhance the reacting surface area, increasing the number of available adsorption sites.
The formation of p–n heterojunctions, as tentatively shown in Fig. 8, improves the gas sensing response of the SnO2-QD/CuO.26,49,50 Taking into account that CuO and SnO2 QDs yield band gaps (BG) of ≈1.8 eV and ≈4.1 eV (Fig. 5b), and assuming that CuO valence (VB) and conduction (CB) bands potentials are higher than those of SnO2,49,50 when the two materials are brought into contact, Fermi levels align and a typical Z-scheme heterojunction is build up.50 Specifically, electrons from SnO2 diffuse into CuO, and holes from CuO migrate into SnO2, forming a depletion region at the interface. Such charge redistribution leads to the establishment of an internal electric field, directed from SnO2 to CuO. As a result, a built-in potential barrier forms, causing an accumulation of electrons in the SnO2 conduction band and holes in the CuO valence band. Consequently, a majority of charge carriers (electrons for SnO2 and holes for CuO) are available to participate in the gas–surface reactions, yielding a stronger modulation of the electrical resistance in the presence of a target gas.
 |
| | Fig. 8 Schematization of the p–n heterojunction at the interface of CuO and SnO2 (a) before and (b) after contact. | |
It turns out that the larger extension of the p-type CuO surface as respect to that of n-type SnO2-QDs, explains the overall p-type response of the SnO2/CuO. Given that, we may also estimate that the CuO surface of the heterojunction is the primarily reacting surface to NO2, H2 and H2O adsorbing molecules. Under these conditions at operating temperatures below 150 °C, the adsorption of oxygen on a doubly positively-charged oxygen vacancy
leads to an increase of the holes h˙ concentration in p-type CuO (R1), resulting in a resistance decrease.
| |
 | (R1) |
NO2 molecules, on the other hand, interacts with CuO by direct adsorption on free vacancy sites (R2).23 Being NO2 a strong electron-acceptor, electrons are released from the surface, leading to an increase of the h˙ concentration in the material, eventually decreasing the resistance (Fig. 6d).
| |
 | (R2) |
When a reducing gas like H2 is introduced, it generally reacts with adsorbed oxygen,53 decreasing h˙ concentration (R3), leading to a resistance increase (Fig. 6e).
| |
 | (R3) |
Regarding water interaction with CuO, a mechanism involving pre-adsorbed oxygen and Cu lattice atoms has been proposed,58,61 which yields a depletion of h˙ concentration (R4) and an increase of the sensor's resistance (Fig. 7b – panel (ii)).
| |
 | (R4) |
These mechanisms also provide a coherent explanation for the synergistic/antisynergistic effects induced by humidity in the presence of oxidizing or reducing gases. Specifically in Fig. 7b and e – panel (ii), we demonstrated that sensor's relative responses (RRs) to NO2/H2 gases in presence of 50% RH increases/decreases as respect to dry air background (see Fig. 7b and e – panel (i)), following a synergistic/antisynergistic response. In case of NO2, the synergistic response in 50% RH can be explained considering that as soon as 50% RH interacts with the sensor's surface (Fig. 7b – panel (ii)), reaction (R4) increases the concentration of
, boosting NO2 adsorption according to reaction (R2). Conversely, the antisynergistic response to H2 in the presence of 50% RH, can be described taking into account that as soon as 50% RH interacts with the sensor's surface (Fig. 7e – panel (ii)), reaction (R4) decreases the concentration of
, therefore hampering H2 reduction according to reaction (R3). DFT atomistic simulations of NO2, H2 gases and H2O molecules adsorption over CuO and SnO2 metal oxides, highlight a substantial agreement with the ionosorption mechanism discussed in this paragraph. Specifically, it was confirmed the stronger oxidizing attitude of NO2 molecules to form NO2− with oxygen-vacancy sites
.62,63 Furthermore, while H2 adsorption on SnO2 is characterized by physisorption,64 its interaction with CuO follows a dissociative adsorption pathway.65 Accordingly, the weak interaction of H2 molecules predicted by DFT calculations, is congruent with the smaller sensor's signal variation to H2 compared to NO2, as illustrated in Fig. 6i. Finally, H2O vapor is reported to chemisorb on SnO2, forming two hydroxyl groups.51 On CuO, H2O adsorbs either chemically at Cu sites or dissociatively at oxygen-deficient regions.66 Both DFT models of H2O adsorption predict a decrease/increase of the electrical resistance for n-type SnO2 or p-type CuO, in agreement with experimental observations.
5. Conclusions
In this work, we have successfully demonstrated a simple and scalable synthesis of a novel heterojunction based on needle-like CuO decorated with SnO2 quantum dots (QDs, d = 3.9 nm) for NO2 and H2 gas sensing applications.
The composite exhibits markedly enhanced gas sensing performance compared to the individual components, with higher sensitivity toward the investigated species and excellent stability and reproducibility of the response. Notably, sensor operates efficiently at 100 °C OT, a significantly lower temperature compared to conventional metal oxide sensors, addressing a key limitation of to date gas sensing technologies. Ultimately, we studied a possible sensing mechanism explaining the role of the heterojunctions and possible gas–surface interaction, tentatively describing also the synergistic/antisynergistic effect played by relative humidity acting as interfering gas over the NO2/H2 responses.
In conclusion, these results point out the improving effect of the CuO and SnO2 QDs coupling, associated to efficient charge transfer and modulation the electronic structure of the sensing interface. Overall, this study provides a promising pathway toward the development of high-performance, low-temperature gas sensors based on engineered oxide.
Conflicts of interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Data availability
The data that support the findings of this study are available from the corresponding author upon reasonable request.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra05533d.
Acknowledgements
C. C. acknowledges the financial support from PRIN 2022 project #2022LZWKAJ CUP E53D23005280006 and from PRIN-PNRR project #P2022JRB2Y CUP E53D23017740001.
References
- S. C. Anenberg, A. Mohegh, D. L. Goldberg, G. H. Kerr, M. Brauer, K. Burkart, P. Hystad, A. Larkin, S. Wozniak and L. Lamsal, Lancet Planetary Health, 2022, 6, e49–e58 CrossRef PubMed.
- P. Huangfu and R. Atkinson, Environ. Int., 2020, 144, 105998 CrossRef CAS.
- T. N. Vezirolu and F. Barbir, Int. J. Hydrogen Energy, 1992, 17, 391–404 CrossRef.
- X. Ma, B. Li, B. Han, X. Jin, Y. Liu, M. S. Bi and C. M. Shu, Int. J. Hydrogen Energy, 2025, 114, 355–367 CrossRef CAS.
- Y. Masuda, Sens. Actuators, B, 2022, 364, 131876 CrossRef CAS.
- M. A. Franco, P. P. Conti, R. S. Andre and D. S. Correa, Sensors and Actuators Reports, 2022, 4, 100100 CrossRef.
- S. Shah, S. Hussain, S. T. U. Din, A. Shahid, J. N. O. Amu-Darko, M. Wang, Y. Tianyan, G. Liu and G. Qiao, J. Environ. Chem. Eng., 2024, 12, 112538 CrossRef CAS.
- O. Baranov, K. Bazaka, T. Belmonte, C. Riccardi, H. E. Roman, M. Mohandas, S. Xu, U. Cvelbar and I. Levchenko, Nanoscale Horiz., 2023, 8, 568–602 RSC.
- N. Kaur, ACS Sens., 2025, 10, 1641–1674 CrossRef CAS PubMed.
- A. V. Almaev, B. O. Kushnarev, E. V. Chernikov, V. A. Novikov, P. M. Korusenko and S. N. Nesov, Superlattices Microstruct., 2021, 151, 106835 CrossRef CAS.
- N. Barsan and U. Weimar, J. Electroceram., 2001, 143–167 CrossRef CAS.
- K. G. Krishna, S. Parne, N. Pothukanuri, V. Kathirvelu, S. Gandi and D. Joshi, Sens. Actuators, A, 2022, 341, 113578 CrossRef CAS.
- C. Cantalini, L. Lozzi, M. Passacantando and S. Santucci, IEEE Sens. J., 2003, 3, 171–179 CrossRef CAS.
- C. Li, X. Chen, Z. Zhang, X. Wu, T. Yu, R. Bie, D. Yang, Y. Yao, Z. Wang and L. Sun, Nano Lett., 2024, 24, 15025–15034 CrossRef CAS PubMed.
- F. Zhou, Q. Xin, Y. Fu, Z. Hua, Y. Dong, M. Ran, H. Song, S. Liu, R. Qu, Y. Yang, X. Zhang, C. Zheng and X. Gao, Chem. Eng. J., 2023, 464, 142471 CrossRef CAS.
- F. J. Meng, R. F. Xin and S. X. Li, Materials, 2023, 16, 263 CrossRef.
- M. Mathew, P. V. Shinde, R. Samal and C. S. Rout, J. Mater. Sci., 2021, 56, 9575–9604 CrossRef CAS.
- Q. Zhang, K. Zhang, D. Xu, G. Yang, H. Huang, F. Nie, C. Liu and S. Yang, Prog. Mater. Sci., 2014, 60, 208–337 CrossRef CAS.
- S. W. Choi, A. Katoch, J. Zhang and S. S. Kim, Sens. Actuators, B, 2013, 176, 585–591 CrossRef CAS.
- A. Govind, P. Bharathi, M. K. Mohan, J. Archana, S. Harish and M. Navaneethan, J. Environ. Chem. Eng., 2023, 11, 110056 CrossRef CAS.
- N. Taguchi, US Pat., 3631436, 1970 Search PubMed.
- D. Gautam, Y. K. Gautam, K. Sharma, A. Kumar, A. Kumar, V. Srivastava and B. P. Singh, Int. J. Hydrogen Energy, 2024, 81, 313–345 CrossRef CAS.
- V. Paolucci, J. De Santis, L. Lozzi, G. Giorgi and C. Cantalini, Sens. Actuators, B, 2022, 350, 130890 CrossRef CAS.
- S. B. Dhage, V. L. Patil, P. S. Patil, J. Ryu, D. R. Patil and Y. S. Malghe, Mater. Lett., 2021, 305, 130831 CrossRef CAS.
- H. Ullah, F. Usman, M. A. Rasheed, A. A. Khan, R. Ahmad, S. Y. Lee, J. Heo, G. Ali, M. Khan and S. O. Cho, ACS Appl. Nano Mater., 2022, 5, 9096–9111 CrossRef CAS.
- G. Charrada, M. Ajili, N. Jebbari, M. Hajji, S. Bernardini, K. Aguir and N. Turki Kamoun, Mater. Lett., 2024, 367, 136666 CrossRef CAS.
- J. Zhang, S. Ma, B. Wang and S. Pei, Sens. Actuators, A, 2021, 332, 113090 CrossRef CAS.
- S. W. Choi, A. Katoch, J. Zhang and S. S. Kim, Sens. Actuators, B, 2013, 176, 585–591 CrossRef CAS.
- S. Li, X. Li, J. Guo, X. Liu and J. Zhang, ACS Appl. Nano Mater., 2025, 8, 10549–10558 CrossRef CAS.
- Z. Song, S. Xu, J. Liu, Z. Hu, N. Gao, J. Zhang, F. Yi, G. Zhang, S. Jiang and H. Liu, Sens. Actuators, B, 2018, 271, 147–156 CrossRef CAS.
- L. Zhu, M. Wang, T. Kwan Lam, C. Zhang, H. Du, B. Li and Y. Yao, Sens. Actuators, B, 2016, 236, 646–653 CrossRef CAS.
- J. Du, R. Zhao, Y. Xie and J. Li, Appl. Surf. Sci., 2015, 346, 256–262 CrossRef CAS.
- Z. Tian, H. Bai, Y. Li, W. Liu, J. Li, Q. Kong and G. Xi, ChemistryOpen, 2020, 9, 80–86 CrossRef CAS PubMed.
- B. K. Sahu and A. Das, J. Phys. Chem. Solids, 2019, 129, 293–297 CrossRef CAS.
- V. Y. Yevdokymenko, O. Dobrozhan, R. Pshenychnyi, A. Opanasyuk, Y. Gnatenko, A. Bukivskii, P. Bukivskij, R. Gamernyk, O. Klymov, V. Muñoz-Sanjosé, P. Ibañez-Romero and Z. Gacevic, Mater. Sci. Semicond. Process., 2023, 161, 107472 CrossRef CAS.
- A. D. Yoffe, Adv. Phys., 2001, 50, 1–208 CrossRef CAS.
- P. Kubelka and F. Munk, Zeitschrift für Technische Physik, 1931, 12, 593–609 Search PubMed.
- S. Saif, S. F. Adil, M. Khan, M. R. Hatshan, M. Khan and F. Bashir, Sustainability, 2021, 13, 2017 CrossRef CAS.
- M. K. Singh, M. C. Mathpal and A. Agarwal, Chem. Phys. Lett., 2012, 536, 87–91 CrossRef CAS.
- R. Kumar, L. Mamta, R. Kumari and V. N. Singh, Micromachines, 2023, 14, 728 CrossRef PubMed.
- P. Makuła, M. Pacia and W. Macyk, J. Phys. Chem. Lett., 2018, 9, 6814–6817 CrossRef.
- M. A. Khan, N. Nayan, S. Shadiullah, M. K. Ahmad and C. F. Soon, Nanomaterials, 2020, 10, 1298 CrossRef CAS.
- L. Brus, J. Phys. Chem., 1986, 90, 2555–2560 CrossRef CAS.
- K. J. Button, C. G. Fonstad and W. Dreybrodt, Phys. Rev. B: Condens. Matter Mater. Phys., 1971, 4, 4539–4542 CrossRef.
- J. P. Mendez and D. Mamaluy, Sci. Rep., 2022, 12(1), 16397 CrossRef CAS PubMed.
- A. Walsh and K. T. Butler, Acc. Chem. Res., 2014, 47, 364–372 CrossRef CAS PubMed.
- J. Berkowitz, W. A. Chupka and D. Gutman, J. Chem. Phys., 1971, 55, 2733–2745 CrossRef CAS.
- S. Steinhauer, Chemosensors, 2021, 9, 51 CrossRef CAS.
- A. Paleczek, B. Szafraniak, Ł. Fuśnik, A. Brudnik, D. Grochala, S. Kluska, M. Jurzecka-Szymacha, E. Maciak, P. Kałużyński and A. Rydosz, Sensors, 2021, 21, 4387 CrossRef CAS PubMed.
- W. dong Zhou, D. Dastan, X. tao Yin, S. Nie, S. Wu, Q. Wang and J. Li, J. Mater. Sci.: Mater. Electron., 2020, 31, 18412–18426 CrossRef.
- V. Paolucci, J. De Santis, V. Ricci, L. Lozzi, G. Giorgi and C. Cantalini, ACS Sens., 2022, 7, 2058–2068 CrossRef CAS.
- H. Bai, H. Guo, C. Feng, J. Wang, B. Liu, Z. Xie, F. Guo, D. Chen, R. Zhang and Y. Zheng, J. Mater. Chem. C, 2022, 10, 3756–3769 RSC.
- S. De, N. Venkataramani, S. Prasad, R. O. Dusane, L. Presmanes, Y. Thimont, P. Tailhades, V. Baco-Carles, C. Bonningue, S. T. Pisharam and A. Barnabe, IEEE Sens. J., 2018, 18, 6937–6945 CAS.
- H. Kheel, G. J. Sun, J. K. Lee, A. Mirzaei, S. Choi and C. Lee, Met. Mater. Int., 2017, 23, 214–219 CrossRef CAS.
- S. Shah, S. Hussain, L. A. Khan, K. Yusuf, R. K. Manavalan, Y. Tianyan, X. Zhang, G. Liu and G. Qiao, Mater. Res. Bull., 2024, 180, 113021 CrossRef CAS.
- S. Nie, J. Li, Y. He and X. Yin, ACS Sens., 2024, 9, 6390–6399 CrossRef CAS PubMed.
- O. Lupan, D. Santos-Carballal, N. Ababii, N. Magariu, S. Hansen, A. Vahl, L. Zimoch, M. Hoppe, T. Pauporté, V. Galstyan, V. Sontea, L. Chow, F. Faupel, R. Adelung, N. H. de Leeuw and E. Comini, ACS Appl. Mater. Interfaces, 2021, 13, 32363–32380 CrossRef CAS PubMed.
- M. Hübner, C. E. Simion, A. Tomescu-Stǎnoiu, S. Pokhrel, N. Bârsan and U. Weimar, Sens. Actuators, B, 2011, 153, 347–353 CrossRef.
- V. Paolucci, D. Mastrippolito, V. Ricci, H. Świątek, T. Klimczuk, L. Ottaviano and C. Cantalini, ACS Appl. Nano Mater., 2024, 7, 3679–3690 CrossRef CAS.
- M. Pelino, C. Cantalini and M. Faccio, Act. Passive Electron. Compon., 1994, 16, 69–87 CrossRef.
- M. Pelino, C. Colella, C. Cantalini, M. Faccio, G. Ferri and A. D'Amico, Sens. Actuators, B, 1992, 7, 464–469 CrossRef CAS.
- J. Su, X. Liu, H. Zhang, B. Zhao, Y. Feng, C. Wang and T. Shen, Phys. Status Solidi B, 2024, 5, 261 Search PubMed.
- S. Kucharski and C. Blackman, Chemosensors, 2021, 9, 270 CrossRef CAS.
- A. Umar, H. Y. Ammar, R. Kumar, T. Almas, A. A. Ibrahim, M. S. AlAssiri, M. Abaker and S. Baskoutas, Int. J. Hydrogen Energy, 2020, 45, 26388–26401 CrossRef CAS.
- G. Hao, R. Zhang, J. Li, B. Wang and Q. Zhao, Comput. Mater. Sci., 2016, 122, 191–200 CrossRef CAS.
- F. Ahmad, M. K. Agusta, R. Maezono and H. K. Dipojono, J. Phys.: Condens. Matter, 2020, 32, 045001 CrossRef CAS PubMed.
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.