Hossein Mojtabazadeh and
Javad Safaei-Ghomi
*
Department of Organic Chemistry, Faculty of Chemistry, University of Kashan, P.O. Box 87317-51167, Kashan, I. R. Iran. E-mail: safaei@kashanu.ac.ir; Fax: +98-31-55552935; Tel: +98-31-55912385
First published on 22nd April 2025
Radio frequency identification (RFID) technology has made significant strides in recent years, opening up a world of possibilities for various industries. However, to achieve success, reliable and accurate real-time data is crucial. One exciting application of RFID technology is fast and wireless detection of gases. Hydrogen, in particular, is considered a clean fuel. However, it is highly flammable, and detecting it quickly and accurately is challenging in various industries. In this regard, our research focuses on developing a high-conductivity graphite paste for RFID tags integrated with a wireless hydrogen sensor based on nano-CeO2–Fe2O3–graphene oxide. In this work, we obtained a graphite paste using Ultra High Power (UHP) graphite electrodes with a high conductivity of 4.75 × 105 S cm−1 for non-metallic substrates and 4 × 106 S cm−1 with aluminum substrate. Furthermore, we incorporated a hydrogen gas detection sensor into the RFID tag utilizing graphene oxide and cerium oxide–iron oxide nanoparticles. The sensor demonstrated high sensitivity to low concentrations of H2 gas (1 ppm), with stable and repeatable performance. The wireless sensing response was evaluated through reflection coefficient (S11) measurements, confirming effective impedance matching between the RFID chip and antenna. Through this research, we aim to promote the advancement of RFID technology by introducing a low-cost, battery-free sensing platform using graphite and nano-engineered materials, suitable for diverse industrial applications.
Graphite electrodes made from petroleum coke and needle coke are a source of high conductivity and low resistance graphite. These electrodes can be categorized based on their power rate as regular power, high power, and ultra-high power (UHP). UHP graphite electrode is an ultra-strong graphite-based conductive material that uses petroleum coke and needle coke as aggregates, coal asphalt as a binder, and undergoes complex processes such as calcining, crushing, grinding, batching, mixing, shaping, roasting, impregnation, graphitization, and mechanical processing. It finds application in industries that require ultra-high power steelmaking electric arc furnaces due to its lower resistance, better electrical conductivity, and higher mechanical strength.14–16 In this research, we intend to prepare an RFID tag based on UHP graphite electrode with high conductivity and low cost.
Hydrogen (H2) is an eco-friendly and unpolluted fuel that does not emit carbon dioxide (CO2) when burned. It can be effortlessly generated by electrolyzing water. H2 is a highly flammable, odorless and colorless gas, hence detecting H2 gas leaks is crucial.17–19 Different types of sensors have been created for H2 detection, such as thermal conductivity detectors, electrochemical sensors, semiconducting metal oxides (SMOX),20 and catalytic combustion sensors.21 SMOX sensors including SnO2,22 ZnO,23 Fe2O3,24 TiO2,25 WO3,26 In2O3 (ref. 27) and CeO2 (ref. 28) have been developed for hydrogen detection. However, the use of CeO2 for H2 detection has not been explored much. CeO2 has two stable oxidation states (Ce4+ and Ce3+) that enable oxygen storage and release. It can transform into a non-stoichiometric oxide without damaging its crystal structure when exposed to high temperatures. Its high thermal stability, oxygen storage capacity,29 and rich oxygen vacancies make CeO2 ideal for gas detection systems. When combined with Fe2O3, a mixed oxide is formed that has higher reducibility at low temperatures than pure CeO2.30–32
The research presents the development of a passive UHF-RFID tag with a wireless H2 gas detection smart sensor. The developed passive UHF-RFID tag includes a sensing region composed of CeO2–Fe2O3 nanoparticles supported on graphene oxide.
![]() | ||
Fig. 1 (A) General scheme for graphite paste fabrication and RFID tag screen printing (B) fabrication of nano-CeO2–Fe2O3–GO and chemical resistance H2 sensing system with wireless detection. |
In this research, it was attempted to achieve suitable adhesion of the UHP-GE paste to the surface without negatively affecting resistance by decomposing the AC and removing the solvent and AMP at a temperature of 180 °C.
Analysis of the Raman spectrum demonstrated indications of both loss of solvent and additives, as depicted in Fig. 2a. The sharp peak at approximately 1570 cm−1 (G) corresponds to the vibrational mode in the plane that includes sp2 hybridized carbon atoms generating the graphite sheets. The band that emerges around 2700 cm−1 (2D) is an outcome of the double-phonon lattice vibrational process. The number of graphene layers or stacking can be estimated by the intensity ratio of the 2D and G bands. The higher the ratio, the fewer the number of layers. For instance, monolayer graphene has an I2D/IG ratio of 2.35 In sample “a” (UHP-GE), the I2D/IG ratio was 0.3, while in sample “b” (after treatment with additives and solvents and curing at 180 °C), it was 0.4.
X-ray diffraction (XRD) investigation (Fig. 2b) of UHP-GE “a” shows a sharp peak at 26.6°, corresponding to (002). After treatment with additives, solvent, and curing at 180 °C, sample “b” shows a (002) peak with lower intensity compared to “a”. The XRD results were consistent with the graphite standard (JCPDS 96-901-2231). The reduced intensity of the (002) XRD peak in sample “b”, along with the slightly increased I2D/IG ratio in Raman analysis, suggests a mild expansion of interlayer spacing or partial exfoliation of graphite sheets after curing.
Furthermore, the similarity between the spectra and the absence of peaks corresponding to functional groups in the Raman spectrum indicate that AC, ethanol, and AMP were decomposed during the curing process. Fig. 2c1 and c2 are field emission scanning electron microscopy (FESEM) images of samples “a” and “b”. EDS and EDS-MAP (ESI Fig. S1 and S2†) also support the decomposition of AC and evaporation of AMP and solvent components.
The electrical properties of the prepared films were studied, and it was found that the AC significantly increases the resistance. The initial resistance of UHP-GE pastes screened on different substrates was high (109 µΩ m). The boiling point of the AC is 99.5 °C. After applying heat to 100 °C, the resistance of the films decreased to about 1000 µΩ m, indicating the decomposition of the AC. At a temperature of 180 °C, the resistance reached 0.14 µΩ m for the film with an aluminum substrate and 2.1 µΩ m for fire-resistant paper and Kapton foil (Fig. 3a). The initial resistance of UHP-GE powder was 3.5 µΩ m. After paste formulation and curing, the resistance decreased to 0.14 µΩ m (on aluminum), benefiting from both improved conductivity and strong adhesion. Increasing the graphite layer thickness from 25 µm to 35 µm via repeated screen printing increased the resistance to 2.1 µΩ m, indicating that the conductive effect of the aluminum substrate diminishes as the graphite layer becomes thicker. It is clear that in the samples with Kapton foil substrate and fire-resistant paper, the resistance also decreases with the increase in the number of screen printing.
![]() | ||
Fig. 3 (a) The graph displays the resistance values at different temperatures. (b) The three-dimensional radiation pattern of the UHF dipole RFID tag antenna model was calculated using HFSS. (c) The S11 results of the fabricated UHF dipole tag antennae were both calculated and measured. (d) The 915 MHz read range pattern of the antenna with an aluminum substrate was recorded, showing superior performance compared to non-metallic substrates (as depicted in ESI Table S3†). |
The maximum gain of antennas with non-metallic substrates (fire-resistant paper and Kapton foil) was 1 dB, and the maximum gain of the antenna with an aluminum substrate was −2 dB (Fig. 3b). The measured S11 values showed good agreement with the simulated results at 915 MHz, confirming proper impedance matching. The “reflection coefficient” (S11) shows the compatibility of the tag antenna with the RFID chip connected to it. The more negative the value of S11, the higher the efficiency of the antenna. S11 was −28 dB for the aluminum substrate antenna (Fig. 3c) and −13 dB for the other antennas (Table S2†). These results show that the proposed antennas match the impedance of the chip at the frequency of 915 MHz. The integration of the sensing layer did not significantly shift the resonance frequency, indicating that the antenna performance remained stable post-fabrication.
Additionally, the reading range of tag antennas with Kapton foil, fire-resistant paper, and aluminum substrates was about 6, 5.9 and 10 m, respectively (Fig. 3d).
The results of this study demonstrate that the graphite tags prepared from UHP-GE paste perform exceptionally well when compared to graphene and even metallic tags (Table 1). Moreover, they are cost-effective and possess long read ranges that make them a viable alternative to expensive tags. By utilizing these tags, we can achieve greater efficiency and accuracy in various industries, ranging from logistics to healthcare.
No. | Material | Rsa (Ω sq−1) | σb (S m−1) | Read range (m) | Year/ref. |
---|---|---|---|---|---|
a Rs = sheet resistance.b σ = conductivity.c FRP = fire-resistant paper.d KF = Kapton foil.e Al = aluminium. | |||||
1 | Graphene | 5 | 1 × 104 | 2.6 | 2016/36 |
2 | Graphene-nanoflakes | 3.8 | 4.3 × 104 | 4 | 2016/37 |
3 | Few-layer graphene | 3 | 1 × 104 | 11 | 2019/38 |
4 | Graphene | 2.8 | 3.7 × 104 | 9 | 2018/39 |
5 | Graphene | 1.9 | 1.39 × 104 | 5 | 2016/40 |
6 | Graphene nanoplatelets | 0.04 | 4.2 × 105 | N/A | 2018/41 |
7 | Graphene | 0.02 | 1.6 × 106 | 12 | 2021/42 |
8 | UHP-GE@ FRPc or UHP-GE@ KFd | 0.08 | 4.75 × 105 | 6 | This work |
9 | UHP-GE@ Ale | 0.01 | 4 × 106 | 10 | This work |
10 | Al | 0.00289 | 1.38 × 107 | 15.2 | This work |
The XRD analysis has confirmed the formation of nano-CeO2–Fe2O3, which is consistent with the XRD patterns of both CeO2 and Fe2O3, as illustrated in Fig. 4f. The pattern indicates that CeO2 is crystallized in a cubic structure (JCPDS file no. 34-0394), with six main reflections at 28.56° (111), 31.90° (200), 47.40° (220), 56.40° (311), 59.10° (222), 69.72° (400), 76.44° (331) and 78.54° (420). Additionally, Fe2O3 is present in the form of hematite and in a rhombohedral network structure, according to (JCPDS file no. 89-0598). Its diffraction peaks at 24.76° (012), 32.86° (104), 35.42° (110), 40.81° (113), 49.31° (024), 53.86° (116), 57.34° (122), 62.40° (214), 63.96° (300), 71.75° (1010), 75.22° (220).
In the second step, nano-CeO2–Fe2O3 was added to graphene oxide in varying ratios (1 to 10 mmol) to investigate the impact of nanoparticle loading on sensor performance. The amount of CeO2–Fe2O3 plays a critical role in determining the conductivity and sensitivity of the final sensing composite. Ethanol, AC, and AMP were then added to create a paste of nano-CeO2–Fe2O3–GO, which was cured using the UHP-GE paste method.
The EDS (energy dispersive X-ray spectroscopy) results depicted in Fig. 4d and e provide insight into the percentage of elements present in the nano-CeO2–Fe2O3 dispersed within the graphene oxide. These results also demonstrate a uniform dispersion of the elements throughout the sample. As shown in Fig. 4f(b), the characteristic diffraction peak at 12.0° confirms the presence of graphene oxide in the composite. Furthermore, the nano-CeO2–Fe2O3–GO patterns exhibit CeO2 peaks at 28.56° (111) and 47.40° (220), as well as rhomboid peaks at 32.86° (104) and 35.42° (110), indicating the presence of Fe2O3. However, the Fe2O3-related peaks are relatively weak and broad, likely due to the small crystallite size, lower relative content, and their fine dispersion within the GO matrix. Moreover, the post-curing step at 180 °C is not sufficient to induce significant grain growth, contributing to the observed broadening. Overall, these results demonstrate that AC can help to better bond graphene oxide with nanoparticles on different surfaces. The AC is decomposed by heat application and does not affect conductivity or resistance.
Due to the broad scope of this work, which included fabrication of the RFID tag in addition to the sensing layer, systematic variation of the Fe/Ce ratio was not performed. However, this parameter could significantly influence crystallinity and sensor performance and is therefore recommended as a topic for future research.
To evaluate conductivity and nanoparticle distribution, I–V curves were recorded under ambient conditions. As shown in Fig. 5a, all samples exhibited linearity from −0.1 to 0.1 V, confirming ohmic contact behavior. Conductivity improved with increasing nano-CeO2–Fe2O3–GO content.
Sensor behavior was assessed by monitoring resistance changes in nano-CeO2–Fe2O3–GO-1, -5, and -10 tags under different H2 concentrations at 25 °C and 20% RH. Pure H2 gas was used in these tests (Fig. 5b). The resistance increased proportionally with H2 concentration, from 1 to 40 ppm. This trend is attributed to the interaction between adsorbed hydrogen and surface-adsorbed oxygen species, which modulate carrier density and enhance electron depletion effects in the semiconducting metal oxide nanoparticles.
Furthermore, nano-CeO2–Fe2O3–GO-1 failed to respond to H2 concentrations below 10 ppm, likely due to insufficient active surface area and limited adsorption capacity. In contrast, nano-CeO2–Fe2O3–GO-10 exhibited clear detection at 1 ppm, attributed to higher nanoparticle content enhancing surface reactions.
However, CeO2–Fe2O3–GO-10 also exhibited the highest baseline resistance, as observed in Fig. 5b. This is due to the semiconducting nature of CeO2 and Fe2O3 nanoparticles, which disrupt the conductive GO network when present in excess. Increased oxide content reduces percolation pathways and carrier mobility, resulting in higher overall resistance. This phenomenon has been similarly reported in oxide–graphene hybrid sensors in previous studies.3,43
To evaluate the dynamic response of nano-CeO2–Fe2O3–GOs and investigate their repeatability and stability, we exposed the sensors to H2 gas with a concentration of 40 ppm at room temperature (Fig. 5c), which demonstrated excellent stability and repeatability.
In addition, for calibration, the responses obtained from the sensors at different concentrations of H2 gas were recorded and processed (Fig. 5d). At low hydrogen concentrations, nonlinear behavior was observed, but when the hydrogen concentration increased, the behavior of the sensor became linear. Concentrations between 1 and 100 ppm show a linear behavior, which indicates the high sensitivity of the sensors to H2 gas.
This linearity confirms the sensor's effective response to hydrogen even in low ppm ranges and supports the practical utility of the CeO2–Fe2O3–GO-10 sample.
It was discovered that pristine graphene oxide does not respond to H2 gas due to the lack of an active surface to absorb H2 molecules. Furthermore, when AC remains undecomposed, the sensitivity to H2 gas is negligible due to the high resistance of the compound. However, after AC breakdown, the surface resistance decreases and the electrical conductivity improves.
The input impedance of the antenna terminal (Zin) can be expressed using Kirchhoff's law, as shown in eqn (1):
![]() | (1) |
The mutual inductance between the sensor and antenna is denoted by “M”, while the angular frequency is represented by “ω”. The reflection value, S11 parameter, changes in response to H2 gas due to the resistance change of the variable resistor at the resonance frequency. Eqn (2) provides a simplified formula for S11, where Z0 is a fixed value of 50 Ω for all samples.
![]() | (2) |
An increase in H2 concentration alters both the magnitude and phase of Zin, leading to a higher S11 value and a slight shift in the resonant frequency due to modified impedance matching.
In Fig. 6, the observed gradual shift in the S11 curves with increasing H2 concentration is due to a change in the real and imaginary components of the sensor's input impedance (Za). As the sensor's resistance increases (with more H2 adsorption), both the resonance depth and position of the minimum S11 value are affected. This results in a combined effect of magnitude reduction and slight frequency shift in the reflection profile. Such shifts are common in passive RFID sensors where the variable resistance alters the resonance condition of the tag antenna. Since the resonant frequency depends on both the real (R) and imaginary (reactive) parts of impedance, changes in Rs (sensor resistance) indirectly shift the resonance toward higher frequencies.44,45
To evaluate the response of the wireless hydrogen gas sensors, backscattering between an RFID reader and antennas connected to a network analyzer was employed. The analyzer generated frequency signals to activate the tag and collected impedance data within the desired range. Impedance matching between the antenna and the tag reflected the signal, allowing the analyzer to measure the reflection coefficient (S11). Variations in H2 concentration caused changes in the sensor's resistance, which modified the input impedance (Z_in) and led to observable shifts in the S11 signal, both in magnitude and frequency.
In the absence of H2, initial resistance varied based on the amount of nano-CeO2–Fe2O3–GO (GO-1, GO-5, GO-10), producing different S11 values. Among them, GO-10 showed the largest S11 shift, due to its higher nanoparticle content and lower resistance. Increasing H2 concentration gradually raised the circuit impedance, leading to changes in S11 while the resonance frequency remained nearly constant across all tags (Fig. 6a–c).
The alterations in S11 following a 10-second exposure to H2 were recorded and analyzed in relation to gas concentration. As illustrated in Fig. 6d, nano-CeO2–Fe2O3–GO-10 displayed the most significant S11 parameter shift across various concentrations. The CeO2–Fe2O3–GO nano-based sensors exhibit excellent response and recovery time (Fig. S6†), as well as favorable sensitivity when exposed to hydrogen gas after AC degradation. An increase in the ratio of nano-CeO2–Fe2O3 to graphene oxide results in an increase in sensitivity, with nano-CeO2–Fe2O3–GO-10 displaying exceptional sensitivity (1 ppm).
The environmental conditions under which hydrogen gas is measured are very important to obtain accurate and reliable results. Two key factors to consider are temperature and humidity. The ambient temperature was precisely controlled at 25 °C using a thermostat to ensure consistency throughout the tests. Secondly, the humidity level was also carefully controlled. Since humidity can interfere with gas adsorption mechanisms, changes in ambient moisture levels may significantly influence the sensor's response to H2. To investigate the impact of humidity, the sensor was exposed to various H2 concentrations (5, 10, 25, 50 and 100 ppm). Higher humidity levels reduce the surface adsorption of H2 molecules by competing for active sites, leading to a decrease in sensitivity and hence a reduced |S11| response, as shown in Fig. 7a. This observed reduction in |S11| response under high humidity is attributed to competitive adsorption: water molecules occupy active sites on the CeO2–Fe2O3–GO surface, thereby limiting H2 access and reducing the sensor's sensitivity. In practical scenarios, this can be mitigated by sensor surface modification or use of hydrophobic coatings to improve humidity tolerance.46,47
Fig. 7b further confirms a relatively linear decrease in |S11| values with increasing relative humidity. To determine the impact of humidity on the sensor's response at different H2 concentrations, we conducted a linear regression analysis on the collected data under varying humidity levels. The slope of the line was fitted and illustrated in Fig. 7c. Furthermore, we obtained diffusion coefficient curves of the sensor under different humidity levels, as displayed in Fig. 7d. By substituting the moisture content into the quartic polynomial fitting curve, we can compute the sensitivity coefficient.
The mechanism behind SMOX resistance sensing has been extensively studied and is quite intricate. According to prevalent theory, it involves a modification in surface electron depletion that results from the interaction between H2 and surface adsorbents.48
This sensing mechanism is not purely resistive; it involves catalytic redox reactions between H2 and chemisorbed oxygen species (O2−) on the surface of the sensing layer. In this composite, CeO2 provides oxygen storage and redox buffering via Ce4+/Ce3+ transitions, enhancing the response time and stability, while Fe2O3 contributes catalytic activity that accelerates H2 dissociation and reaction kinetics. When nano-CeO2–Fe2O3–GO is exposed to air, O2 molecules adhere to its surface and are transformed into O2− on the surface of the material, as depicted in Fig. 8. This process leads to the formation of an electron layer near the surface, which increases resistance. However, when the surface of nano-CeO2–Fe2O3–GO is exposed to H2, an exothermic oxidation–reduction reaction takes place between H2 and O2−, which rapidly eliminates H2O molecules from the surface. This, in turn, reduces the resistivity of the nano-CeO2–Fe2O3–GOs. Upon returning the sensor to ambient air, the adsorbed species reabsorb the region, causing the resistance to revert to its initial level after the H2 response. Simply put, O2 adsorbs on the surface in air, increasing resistance; upon H2 exposure, it reacts with O2−, releasing electrons and lowering resistance. The process is reversible in air.
The mechanism of hydrogen measurement can be expressed as follows:
O2 (g) + e− → O2− (ads) | (3) |
H2 (g) + 1/2O2− (ads) → H2O + e− | (4) |
A comparison of the sensors prepared in this work with other papers is given in Table 2.
Material | Temperature (°C) | Con.a (ppm) | LODb (ppm) | tresc/trecd | Year/ref. |
---|---|---|---|---|---|
a Con = concentration.b LOD = limit of detection.c res = response time (second).d rec = recovery time (second). | |||||
In2O3/CeO2 | 160 | 50 | 0.01 | 1/9 | 2018/43 |
Pd-NP/CeO2 | 350 | 106 | 10 | 5/17 | 2021/49 |
CeO2/SnO2 | 300 | 60 | 5 | 17/24 | 2018/50 |
Pd@CeO2 | 350 | 100 | 1.4 | 60/360 | 2021/51 |
CeO2-Pd-PDA/rGO | 100 | 6000 | 200 | 70/180 | 2022/52 |
Nano-CeO2–Fe2O3–GO | RT | 40 | 1 | 8/60 | This work |
In order to assess the electrical characteristics and chemical resistance of the H2 sensor, we utilized RFID tag based on nano-CeO2–Fe2O3–GO. The tags were then tested in a sealed chamber (as depicted in Fig. S4†) at room temperature (25 °C). The RFID reader was positioned 15 cm above the tag in the chamber. Following this, the tags were stabilized within the chamber with fresh air flow for a period of two hours. To examine the efficacy of the tags, we exposed them to varying concentrations of H2 gas (ranging from 1–40 ppm), with careful control of gas flow via an MFC (FC-980 MFC Mass Flow Controller 50 SCCM Gas H2). Real-time monitoring of resistance was conducted by applying a constant current of 5–10 amps. After exposure to H2 gas for 5 minutes, free air was introduced into the vacuum chamber for a further 5 minutes. This step served to remove any hydrogen molecules attached to the CeO2–Fe2O3–GO nanoparticles, thereby restoring the active surface of the sensor and allowing it to be reused for repeated measurements of sensor performance. Each tag was read and checked individually. All measurements were made at 4 W EIRP output power.
The formula (R = ΔR/R0 × 100) was used to calculate the “Response” of the sensor to hydrogen gas (R). This formula measures the change in the resistance of the sensor (ΔR) compared to the initial resistance of the sensor (R0). In other words, the response of the sensor is the ratio of the change in resistance (ΔR) to the initial resistance (R0). This formula is commonly employed to gauge a sensor's response to a specific gas.19 All tests are conducted at room temperature (25 °C). Following this, the sensors are stabilized in air for two hours before being tested to check their functionality. Pure hydrogen gas is injected into the chamber for 5 minutes, followed by free air for the next 5 minutes. Throughout the investigation, the sensors are exposed to varying concentrations of H2 gas, ranging from 1–40 ppm. To clarify, we measured the initial resistance of the sensor (R0) in clean air (without H2 gas). When H2 gas was injected into the sensor, the resistance visibly changed, denoted as ΔR. For example, the initial resistance of the sensor (R0) was 0.14 µΩ m (at room temperature and 20% RH). In the presence of 1 ppm of hydrogen gas (in the nano-CeO2–Fe2O3–GO-10), the resistance of the sensor reached 0.14252 µΩ m (ΔR = 0.00252 µΩ m). Using formula (R = ΔR/R0 × 100), we calculated the response of the sensor to hydrogen gas as follows: R = (0.00252 µΩ m)/(0.14 µΩ m) ×100 = 1.8. Therefore, the response of the sensor to hydrogen gas in this example is equal to 1.8 (Fig. S5a†). The sensitivity of the prepared sensors was determined by calculating the response graph in relation to the concentration, as shown in Fig. S5c and d.† Additionally, Fig. S5a† presents the real-time dynamic response of the sensor under repeated hydrogen exposure and recovery cycles, confirming its wireless operation and reversibility.
Footnote |
† Electronic supplementary information (ESI) available: Resistance, EDS, RFID antenna, and H2 sensitivity data. See DOI: https://doi.org/10.1039/d5ra00587f |
This journal is © The Royal Society of Chemistry 2025 |