1. Introduction
Fossil fuel usage has risen dramatically, about eightfold since the 1950s and approximately doubling since the 1980s. Fossil fuels have been fundamental to building industrialized societies worldwide and accumulation wealth. In contrast, their detrimental effects on environment, human health, and the climate suggest that we must move away from them. Despite obviously the worst climate crises across the globe affecting everyday life severely, it is apparent that the energy industry will utilize fossil-based fuels more in the coming decade [
1].
We have witnessed record utilization of wide range of fossil-based energy sources after the industrial revolution. In 2019, the production output of oil and petroleum reached a record high. Similarly, natural gas plant liquids (NGPL) production has been rising steadily since 2005 and reached a record high in 2023. Renewable energy output in 2023 was about 9%. The contribution of coal to total U.S. energy output has declined from 37% in 1950 to 9% in 2023, equalizing it to the total renewable contribution in the same year. The anticipated upsurge in the accumulation of hazardous emissions (H
2S and NO
2) into the environment places greater importance on the environmental and operational concerns associated with hydrocarbon management and energy production-conversion in the near term. The major energy production in the USA by source is as follows; natural gas 36%, oil 38% followed by coal 9%. Fossil fuels (oil, natural gas, and coal) retained 78% of energy production in 2017 which increased to 83% in 2024, while renewable energy is only 9% of the total output [
2,
3,
4].
One of the main drivers of the gas sensor market’s rapid growth is the need to monitor environmentally hazardous gases [
5]. Gas sensors for pollutants and toxic gases such as nitrogen dioxide (NO
2) and hydrogen sulfide (H
2S) receive significant attention to safeguard human health, nature, and habitat [
6,
7]. These sensors are also used to optimize various energy production-conversion, regulate exhaust emissions and desulphurization [
8]. The NO
2 and H
2S are principal air pollutants causing harm to ecosystem balance, destroying the stratosphere’s protective ozone layer, and accelerating deforestation. NO
2 is easily oxidized to nitric acids. H
2S is also infamously known to be used as a chemical warfare agent during WW2 [
9,
10,
11,
12]. The primary and secondary exposure limits for NO
2 is 1 ppm arithmetic mean annual concentration, while it is for H
2S 5 ppm [
13].
NO
2 and H
2S emissions are primarily produced by the combustion of natural gas, oil, and coal in refineries, process furnaces, fluid catalytic cracking (FCC) regenerators, electric power plant boilers, gas turbines where they are converted into serviceable products for energy production-conversion for commercial and residential use [
14,
15,
16]. 65% of total NO
2 emission is produced by human activity stemming from combustion processes such as; cars, ships, airplanes, petroleum refineries, industrial processes, and power plants. Handling H
2S and NO
2 is challenging at every phase of hydrocarbon including production, purifying, and transport. Emission control has become a major challenge due to the increased utilization of fossil fuels by both developing and developed nations [
17,
18,
19,
20,
21,
22].
Solid oxide fuel cells (SOFCs) are promising energy production conversion industry due to their low-carbon footprint without hazardous exhaust emissions such as H
2S and NO
2 [
23]. SOFCs can utilize a variety of renewable, environmentally friendly inexpensive fuel sources such as natural and landfill gases, coal syngas, biomass, sewage, municipal, and green wastes. Considering the composition of the fuels, the emission of the following hazardous gases is anticipated; NO
2 and H
2S.
The Ni-based (nickel oxide) composite is the most common anode material for SOFCs; however, the just-mentioned gases found in fuel sources cause degradation of SOFCs by poisoning the Ni-based anode. As a result, the energy conversion-production industry as well as green energy production initiatives will greatly benefit from real-time monitoring of H
2S and NO
2 concentration to regulate/maintain optimal fuel utilization. In addition to that, safe-guarding environmental protection will be in place without expensive and cumbersome regulations-filtering-testing requirements [
17,
18,
24,
25].
H2S detection is realized by the following MOS-based gas sensors
; WO
3, CeO
2, SnO
2, ZnO, CuO, platinum and palladium oxides, Fe
2O
3, In
2O
3, TiO
2, CdO [
5,
26,
27,
28,
29]. WO
3 and SnO
2 lacked stability and showed cross-sensitivity in addition to a requirement for high temperature (>300°C).
Table 2 provides literature findings for H
2S sensors, materials, and performances together with results obtained in our current work. Liewhiran et
al. used SnO
2 for H
2S sensing at 300°C, the sensor response (S) was 8 for 10 ppm H
2S [
30]. Some mixed-modified unconventional materials were investigated for H
2S at lower temperatures. Fe
2O
3-Fe
2(MoO
4)
3 showed high sensitivity for H
2S at 225°C. La
0.7Pb
0.3Fe
0.4Ni
0.6O
3 and CdIn
2O
4 did not exhibit sufficient sensing at a temperature range of 100-350°C. Varying ferrites were also tried but the magnitude of sensitivity was in the range of 0-10% resistivity change between 100-400°C [
31,
32,
33,
34]. Kersen utilized Fe
2O
3 with Fe
2(MoO
4)
3 thick oxide films for 1, 10 and 20 ppm of H
2S at 225°C [
5,
34]. Sun
et al. utilized carbon nanotube templated hematite (α-Fe
2O
3) against 10 ppm H
2S at 130°C [
29]. Chaudhari
et al. concluded that TiO
2 with 5 wt.% Al
2O
3 and 0.5 wt% Pd increased the sensor response (S) up to 0.8 at 250°C for 200-1000 ppm of H
2S, while beyond 350°C a step decrease was detected in sensor response [
35,
36]. Ciftyurek
et al. attributed this detrimental decrease to the abrupt desorption of chemisorbed oxygen species after 350°C by showing a sudden drop in the amount of chemisorbed oxygen species by using photoelectron techniques [
37,
38]. Shirsta et
al. reported on polyaniline nanowires decorated with gold nano-particles for the detection of H
2S at 25°C [
39].
NO2 detection materialized through mainly ZnO, WO
3, TiO
2, SnO
2, In
2O
3, V
2O
5, and NiO. Numerous synthesis/deposition approaches such as sputtering, spray pyrolysis, atomic layer deposition (ALD), chemical vapor deposition (CVD), physical vapor deposition (PVD), sol-gel, spray pyrolysis, hydrothermal technique have been utilized to produce ZnO sensor in pure or doped-alloyed forms within various multi-component heterostructures for NO
2 sensing Cai et al. used gold (Au) doped ZnO nanowires activated via ultraviolet light (UV) that produced a 2.3 sensor response to 1 ppm of NO
2 25°C [
40]. ZnO equipped with reduced graphene oxide (rGO) detected NO
2 at 110°C [
41]. Doping ZnO with Ni showed 108% increase in electrical resistance against 100 ppm NO
2 at 200°C [
42]. Au-polyaniline/ZnO nano-composite showed a sensor response of 21 against 50 ppm of NO
2 at 300°C [
43].
Table 2 provides literature findings for H
2S and NO
2 sensors, materials and performances together with results obtained in our current work. Mai et al. reported on ALD deposited 50 nm thick ZnO sensor for NO
2 sensing at 330°C without any oxygen interference but the sensor response was fairly low due to the confined-densely packed microstructure of ZnO and lower concentration of oxygen vacancies
() [44]. In other words, this low sensor signal is the outcome of a lack of porosity-tortuosity due to the jam-packed grain structure on the ZnO surface. This densely packed microstructure decreases the both effective ZnO surface area that come in contact with NO
2 and target gas penetration through the thickness of the ZnO.
ZnO thin films have been a dynamic field of research and application due to their applications in electronics, biomedicine, sensing applications, transducers, optoelectronics, and catalysts since the 1960s. ZnO is a direct, wide-band gap metal oxide semiconductor (MOS). The most standard utilization form of ZnO is its polycrystalline form. ZnO has direct band gap energy of 3.37 eV. ZnO has a hexagonal lattice and is made up of two interconnecting sublattices of Zn
2+ and O
2− in wurtzite crystal symmetry such that each O
2- ion is enveloped by tetrahedra of Zn
2+ ions. This architecture’s large piezoelectric and pyroelectric properties result from missing the center of symmetry [
45].
Metal oxide semiconductor (MOS) based sensors are the most frequently used gas sensors, due to their straightforward operation principle with high sensitivity, compatible with inexpensive mass fabrication approaches such as screen printing and MEMS. Conversely, MOS sensors suffer from high energy consumption due to high working temperature which is required for the generation-activation of chemisorbed oxygen species. The oxygen species are essential for gas sensing in metal oxide-based sensors. Adsorption of molecular oxygen via chemisorption and dissociation onto stoichiometric metal oxide surfaces is not possible while both are readily feasible on non-stoichiometric metal oxide surfaces [
37,
38,
46].
In our current work, we selected ZnO due to its easiness to integrate desired surface physical and chemical features at lower temperatures. By design, we are aiming to have proper surface chemistry promoting to accommodate a high amount of chemisorbed oxygen species in addition to desired surface texture-porosity facilitating gas diffusion through porosity network open to the external surface. The creation of more active oxygen species such as (,) is favorably associated with the ZnO surface activity. Oxygen vacancy sites () function as dissociative centers to convert oxygen molecules into to desired oxygen species (,).
We remarkably improved the sensor response of ZnO at lower temperatures using defect chemistry and sintering principles. By creating an open microstructure for gas diffusion and oxygen vacancy abundance on the ZnO surface, we were able to increase the effective surface area and tortuosity of ZnO and enhance the catalytic sensing reactions required for sensing to be realized at lower temperatures. In this work, we created oxygen vacancy-abundant high porosity ZnO by heat treatment leading to controlled coarsening of ZnO nano-sized grains under low temperature sintering conditions applied.
The gas sensing mechanism and sequence were explained based on electrical and surface chemistry measurements. The microstructural characterization was completed through SEM analysis. Surface chemistry of ZnO and stoichiometry analysis for the oxygen vacancy () concentration were realized through AES, XPS, and UPS measurements. NO2 detecting in combustion environments such as an automobile exhaust is challenging because of coexistence of O2 and NO2 in combustion processes, this is because we utilized oxygen (O2) gas together with NO2 throughout the sensor tests to realize realistic testing conditions.
The H2S and NO2 detection using MOS at low temperatures with high sensor response without an expensive catalyst and/or other cumbersome modification techniques is very challenging. In our work, we developed Engineered Porosity approach leading ZnO with high porosity and tortuosity together with abundant oxygen vacancies () that enabled NO2 and H2S sensing at 200°C with an extraordinarily high sensor response. We developed a micro gas sensor architecture based on Engineered Porosity ZnO with high tortuosity and abundant oxygen vacancies.
3. Result
3.1. Conception and Creation of Engineered Porosity ZnO with Oxygen Vacancies ()
ALD makes use of low deposition temperature for metal oxides resulting in fine grain size with low surface roughness and limited porosity which all severely reduces surface area. Because of their limited porosity, metal oxides produced by the ALD technique have hardly any gas adsorption sites in as deposited state. From the perspective of gas sensor designer adsorption of oxygen species (,) on MOS is crucial for gas sensing action. The sensing performance of MOS sensor can be improved by generating more gas adsorption sites, creating more oxygen vacancies (), and boosting the surface catalytic activity.
These objectives can be accomplished in a variety of ways, including surface alloying, nano-composite formation, ion bombardment, precious metals additives, and nano-size texturing. In order to create limited grain coarsening and grain growth on the ZnO surface that increases the open porosity network, surface area, and vacancy abundance, we used a unique heat treatment technique in our work.
Grain-to-grain contact and coarsening can reduce the total surface area and minimize the thin film’s total interfacial and surface energy [
18,
47,
48,
49,
50,
51,
52,
53,
54,
55]. This work will primarily concentrate on the development of coarsening and grain growth through the distinct control of a sufficient grain-to-grain to contact with high-level porosity network through the ZnO film thickness. We created
Engineered Porosity ZnO enriched with oxygen vacancies to have inter-grain contacts without extensive necking thus having a large porosity network with vast tortuosity. Fine homogeneously distributed and interconnected porosity established throughout the material which is crucial for target gas distribution pathways and essential to expanding the contact area between sensing material ZnO and porosity interface.
3.2. ZnO Microstructure Analysis
ZnO was deposited to a thickness of ~30 nm, with a grain size of ~15 nm.
Figure 2-a shows the as deposited state of the ZnO; a poorly percolated ZnO microstructure maintains low tortuosity throughout the polycrystalline thin film. Low-level grain coalescence is the result of granular microstructure produced by low-temperature sputtering and evaporation thin film deposition techniques. The thin films having nano-size grains maintain high driving force for grain growth and coarsening.
As shown in
Figure 2-b, a vermicular-shaped ZnO microstructure with an average grain size of ~100 nm was produced in our work following 4 hours of annealing at 300ºC. The SEM micrograph also demonstrates how the annealing of ZnO formed a structure that included more porosity, tortuosity and clearly visible ZnO grains. In contrast to materials with larger grain sizes, the thermodynamic stabilization of surface area minimization is the driving force behind the unique features of grain growth/coarsening during the sintering of nano grain-sized ZnO.
High-level of gas adsorption sites and a fine and homogeneous granular structure are achieved. Heat treatment that resulted in grain coarsening greatly expanded the gas diffusion pathways. It is crucial to note that ZnO coarsening resulted from grain boundary migration, and that the grain boundaries did not impede the grain coarsening because they also co-occurred with the reduction-induced creation of oxygen vacancies.
Figure 2.
30 nm thick ZnO; SEM micrograph (a) before heat treatment (b) after heat treatment at 300°C for 4 hours.
Figure 2.
30 nm thick ZnO; SEM micrograph (a) before heat treatment (b) after heat treatment at 300°C for 4 hours.
A high degree of tortuosity resulted from the controlled sintering/coarsening process, which simultaneously increased porosity and grain boundaries.
Section 3 will go into detail about the chemistry of gas adsorption sites. The heat treatment led to the abundance of oxygen vacancy concentration on the surface, together with well-defined gas diffusion pathways with high tortuosity. Both produce ideally perfect gas sensing conditions.
3.3. Zinc oxide (ZnO) Electrical Resistance Investigation
Normally, ZnO maintains a small amount of oxygen deficiency. This makes ZnO transition metal oxide oxygen deficient n-type semiconductor with a typical electrical resistivity of 0.01-3 Ω•m at room temperature [
58,
59,
60,
61]. The substantial deviances among the reported electrical resistivity values in the literature are linked to diverse deposition systems and/or post-deposition treatment resulting in different microstructures, stoichiometries, and surface properties. Oxygen deficiency, in other words, oxygen vacancies has a strong effect on the electrical, chemical and sensory properties of the ZnO. The oxygen vacancies and oxygen deficiency were investigated and examined using UPS and XPS in the next two sections.
Figure 3.
Electrical resistivity measurement for 30 nm thick ZnO film from 25°C to 350°C.
Figure 3.
Electrical resistivity measurement for 30 nm thick ZnO film from 25°C to 350°C.
Error! Reference source not found. displays the resistivity values of the 50 nm thick ZnO thin film. The electrical resistivity as a function of temperature was measured under environmental conditions in the 25-350°C temperature range. An increase in temperature continually reduced the resistivity values for ZnO thin film. The resistivity at room temperature (25°C) was 1.45 Ω•m, and it showed a sharp decrease, down to 0.30 Ω•m at 140°C. When the temperature was raised to 150°C, the resistivity decrease rhythm slowed. Due to lattice uptake of adsorbed oxygen, some of the oxygen vacancies ( on the surface were annealed, causing a slight decrease after 200°C.
3.4. ZnO Valence Band Investigation
The Zn 4s valance electrons form a bond with the oxygen (O) 2p valance electrons in ZnO, where Zn ions are in tetrahedral coordination with O ions. The bonding between Zn and O ions is ionic, owing to the large variance in their respective electronegativity; as Zn possesses 1.65 while O has 3.44. We measured the ZnO valance band with 150 and 300 eV excitation X-ray photons to probe the surface (0.5 nm depth from the surface) and subsurface areas (1 nm depth from the surface) independently.
Figure 4 shows the valance band region of ZnO in the vicinity of the Zn 3d peak.
Figure 4-a shows Zn 3d and valance band region together with distinct bordering. The valence band is represented by Zn 4s-O 2p blended states near 3-8.5 eV while Zn 3d is observable near 10.7 eV. The broad features of the Zn 4s-O 2p valance band suggest that surface stoichiometry is highly influenced by the high amount of oxygen vacancies (
.
At 1 nm depth measurement, the valance band shows defect-free features of fully stoichiometric ZnO. A decrease in the occupation of the Zn 4s+O 2p blended states from 1 nm to 0.5 nm depth measurements indicates an increase in the concentration of surface oxygen vacancies. At 0.5 nm measurement depth; oxygen vacancies (were observed in the valance region as they gave rise to a perturbation around 7-8 eV at the same time causing a decrease in the intensity around 4-7 eV region.
As seen in
Figure 4-b, the deeper measurement (1 nm depth, black colored) confirmed the stoichiometric ZnO phase. However, 0.5 nm (red colored) depth measurement showed the dominant existence of oxygen vacancy (
rich ZnO
x phase on the surface. The intensity of Zn 4s
+O 2p blended state decreased from ZnO to ZnO
x thus showing the outer surface is dominated by reduced-oxygen deficient-oxygen vacancy rich ZnO
x phase.
Figure 4.
Engineered Porosity ZnO; 1 nm and 0.5 nm depths from the surface (a) Valance band and Zn 3d regions together (b) Enlarged view of the valence band region; Zn 4s+O 2p blending feature.
Figure 4.
Engineered Porosity ZnO; 1 nm and 0.5 nm depths from the surface (a) Valance band and Zn 3d regions together (b) Enlarged view of the valence band region; Zn 4s+O 2p blending feature.
3.5. ZnO Auger Peak (AES) Analysis for Stoichiometric Understanding
AES has larger chemical shifts than XPS core-level shifts, so it allows chemical state analysis in cases that are unlikely for XPS [
59]. Determination of the oxidation state of zinc (Zn) will contain significant uncertainties if it is solely based on XPS analysis of zinc (Zn) 2p photoelectron line. In the best scenario, Zn 2p demonstrates a minor shift of ~0.2 eV in the binding energy between zinc (Zn) metal and fully oxidized ZnO; in most of the cases, there is a strong overlap [
62]. The binding energy for Zn 2p
3/2 for Zn metal and ZnO are 1021.7 eV and 1021.9 eV, respectively [
63,
64]. Due to this slight shift in fully oxidized ZnO and Zn metal; perform quantitative-qualitative peak analysis and determination of the oxidation state of Zn based XPS 2p analysis is very challenging. To overcome this obstacle; we proposed to utilize AES in conjunction with XPS.
In situations where XPS is unable to identify the chemical state, a shift of about 4 eV in the Auger peaks distinguishes Zn metal from ZnO. AES is an effective surface sensitive technique capable of probing down to 1 nm depth below the surface. Using the Auger line L
3M
45M
45, a precise chemical state analysis of ZnO was performed. AES peaks are historically represented in the electron kinetic energy scale [
65].
Figure 2 shows L
3M
45M
45 position for ZnO. According to published reports in the literature, L
3M
45M
45 values for in ZnO range from 987 eV to 989 eV. In contrast, L
3M
45M
45 for metallic Zn ranges from 992 eV to 993 eV.
Table 1 provides the bulk of the XSP and AES reported literature along with our findings for ZnO.
Figure 5.
ZnO L3M45M45 position with 1147 eV excitation x-ray photons.
Figure 5.
ZnO L3M45M45 position with 1147 eV excitation x-ray photons.
We report 987.9 eV for zinc oxide (ZnO) L
3M
45M
45. On the other hand, in the
Figure 2, we propose that the shoulder seen on the higher kinetic energy site located at 992.3 eV presents ZnO
x. ZnO
x phase contains large amounts of oxygen vacancies and this is because it showed up in the similar region that metal Zn reported in the literature. Thanks to a larger kinetic energy separation between ZnO and ZnO
x; the shoulder strongly confirmed the existence of oxygen vacancy (
rich zinc oxide (ZnO
x) as a minority phase within the measurement depth of 1 nm. The amount of the corresponding phases will be calculated from the XPS measurements presented in the following section. We showed that the outer ZnO surface is covered by oxygen vacancy rich ZnO
x phase.
Table 1.
Core Level XPS and AES Literature findings for ZnO in comparison with our measurement results.
Table 1.
Core Level XPS and AES Literature findings for ZnO in comparison with our measurement results.
Electron Level |
Zn 3s in ZnO |
Zn 3p1/2 in ZnO |
Zn 3p3/2 in ZnO |
Zn 3d in ZnO |
O 1s in ZnO |
L3M4,5M4,5 in ZnO |
L3M4,5M4,5 in Zn |
Ciftyurek and Schierbaum (this work) |
140.0 |
91.6 |
88.8 |
10.7 |
529.8 |
987.9 |
992.3 |
Vesely and Langer [66] |
139.8 |
92.0 |
89.0 |
10.5 |
530.9 |
988.9 |
- |
Gaarenstroom and Winograd [67] |
- |
- |
- |
10.7 |
- |
987.7 |
- |
Kowalczyk [68] |
- |
- |
- |
- |
- |
- |
991.9 |
Wagner [69] |
- |
- |
- |
- |
- |
- |
992.0 |
Schoen [65] |
139.6 |
91.8 |
88.7 |
10.3 |
530.3 |
988.5 |
992.5 |
Barr and Hackenberg [70] |
- |
- |
- |
10.3 |
530.3 |
987.9 |
992.0 |
Klein and Hercules [71] |
- |
- |
- |
10.4 |
- |
988.2 |
992.3 |
Strohmeier and Hercules [72] |
139.2 |
- |
88.3 |
- |
529.9 |
988.9 |
992.4 |
Powell [73] |
- |
- |
- |
- |
- |
- |
992.4 |
Ley and Kowalczyk [74] |
- |
- |
- |
10.4 |
- |
- |
991.9 |
Wehner and Mercer [75] |
- |
- |
- |
- |
- |
988.1 |
992.1 |
Dake and Baer [76] |
- |
- |
- |
- |
- |
988.1 |
992.2 |
3.6. Concentration of Oxygen Vacancy ( in Engineered Porosity ZnO
Oxygen vacancies facilitate replenishment of the adsorbed oxygen (,) on the ZnO sensor surface which is essential for chemical sensing in metal oxides. Therefore, it is important to measure the concentration of the oxygen vacancies (. Concentration of oxygen vacancies can be calculated either from the oxidation state of the zinc (Zn) in the ZnO or oxygen (O) 1s peak analysis. In our current work, we both analyzed Zn and O via XPS in Engineered Porosity ZnO enriched with oxygen vacancies.
3.6.1. Zn 3p Analysis
Auger analysis proved existence of ZnO
x phase on the ZnO surface. We continued with Zn 3p peak analysis to determine the amounts of each ZnO and ZnO
x. The Zn 3p spectrum from the depth of 1 nm is shown in
Figure 4. Deconvolution analysis was used to quantify the amounts of ZnO and oxygen vacancy (
abundant non-stoichiometric ZnO
x. The Zn 3p envelope was deconvoluted into two distinct chemical states; stoichiometric ZnO and ZnO
x. We measured binding energy of 3p
3/2 for ZnO
x is 87.4 eV while 3p
3/2 for ZnO is 88.8 eV. The binding energies (BE) we reported for 3p
3/2 for ZnO and ZnO
x are in good agreement with literature values [
65,
66,
72,
77]. The decrease in the 3p electron binding energy in ZnO
x in comparison to the ZnO is due to high level of oxygen vacancies created in the ZnO
x as it was processed through heat treatment-sintering-coarsening for creation porosity and tortuosity.
The amount ZnO and ZnOx phases are 33 at.% and 67 at.%, respectively. ZnOx being a dominant phase and homogeneously distributed through the surface and will dictate the sensors interaction with target gaseous environment.
Figure 6.
Deconvoluted Zn 3p peak, showing non-stoichiometric ZnOx associated with oxygen vacancies.
Figure 6.
Deconvoluted Zn 3p peak, showing non-stoichiometric ZnOx associated with oxygen vacancies.
3.6.2. Oxygen 1s Analysis
Figure 7 shows the XPS spectrum of oxygen O 1s spectrum from
Engineered Porosity ZnO surface. The quantification of different oxygen ion containing species; water/hydroxide groups (H
2O/OH
−), chemisorbed oxygen ions
(), and lattice oxygen ions connected to ZnO
x and ZnO were completed. Analysis also focused on the calculation amounts of ZnO and ZnO
x separately. The O 1s spectra are fitted to four sub-spectra centered at 528.98, 529.80, 530.80, and 532.10 eV, accounting for the lattice oxygen ions in ZnO, lattice oxygen ions in oxygen vacancy rich ZnO
x, chemisorbed oxygen ions and water-hydroxyl groups, respectively. The binding energies we report for water/hydroxide, chemisorbed oxygen, and lattice oxygen ions are in good match with literature reported for ZnO and other metal oxides used for gas sensing [
18,
37,
38,
78,
79].
Amount of chemisorbed oxygen ions and water-hydroxyl groups are 30 at.%, and 5 at.%, respectively. The high amount of chemisorbed oxygen ions shows that ZnOx is rich in oxygen vacancies since the chemisorbed oxygen ions attach oxygen vacancy ( and ) centers on ZnO surface. The amount of lattice oxygen ions connected to ZnO is 22 at.%. However, the amount of lattice oxygen ions in oxygen vacancy rich ZnOx is 43 at.%. That shows that majority of the ZnO sensor surface interactions is dictated by the oxygen vacancy abundant ZnOx.
Figure 7.
O 1s spectrum with 545 eV excitation photon energies corresponding to 1 nm depth from the Engineered Porosity ZnO sensor surface. Binding energies and amount of each component are also provided in the graph.
Figure 7.
O 1s spectrum with 545 eV excitation photon energies corresponding to 1 nm depth from the Engineered Porosity ZnO sensor surface. Binding energies and amount of each component are also provided in the graph.
3.6.3. Zinc (Zn) to Oxygen (O) Stoichiometric Quantification (ZnO)
We used intensity factors and ratio between the normalized areas of the O 1s and Zn 3p contributions, O/Zn was found to be 0.73 at.%. We determined x in ZnOx, x=0.73, so the oxygen vacancy abundant phase is ZnO0.73. That shows that majority of the ZnO sensor surface is governed by the oxygen vacancy abundant ZnO0.73.
3.7. Gas Sensor Testing for H2S and NO2
3.7.1. An overview of the sensing mechanism and the involvement of oxygen ions in metal oxide semiconductor (MOS) sensors’ gas sensing reactions
The sensing mechanism for all MOS sensors explained on the basis of adsorption/chemisorption of oxygen ions on MOS surface. Concentration of a certain specific type negatively charged adsorbed/chemisorbed oxygen ions dictates the numbers of electrons in MOS sensor conduction band thus establishes a constant electrical resistance (R).
are the most common chemisorbed oxygen species taking place in gas sensing reactions [
37,
38,
80,
81,
82,
83,
84,
85].
Equations (2) and (3) illustrate adsorption of oxygen gas and consumption of conduction band electrons from ZnO leading to an increase in electrical resistance (R). Subsequently, adsorbed oxygen ions dissociated on the ZnO surface through the processes governed by Equation 4 through Equation 6, requiring consumption of additional ZnO electrons and raising the electrical resistance even further. Equation 7 shows the annihilation of the oxygen vacancy site via adsorption and dissociation of oxygen molecules into vacancy location under suitable temperature and oxygen partial pressure conditions. Target gases such as H
2S, CO, SO
2, etc. consume the adsorbed oxygen ions thus leading to return of electrons to ZnO conduction band, decreasing the electrical resistance [
18,
48,
80,
81,
82,
84,
85,
86]. Equations presented from Equations (2) to (7) are reversible if the certain conditions are satisfied.
Monoatomic oxygen ions (
) are compared to molecular oxygen ions (
) [
37,
38]. The creation of
(Equation 6) is an exothermic reaction that results in ~1 eV decrease in free energy, whereas other reactions presented Equation 1 through 7 are endothermic.
transformation to
occurs at ~150–200°C (Equations 3 and 5). The dissociation of O
2 into
(Equation 5) necessitates ~0.5 eV. Monoatomic oxygen ions
() can tolerate temperatures up to >400°C before departing the ZnO surface.
ions comparing to
, and
are more long-lasting at elevated temperatures on a ZnO surface, , moreover if
ions are trapped (Equation7) by
and
vacancy centers,
ions cannot be distinguished from the lattice oxygen ions. The creation of
(Equation 4) requires about 5 eV, and a formation of
(Equation 6) will require higher energy of ~20 eV [
37,
38,
81,
82,
87,
88].
The most active and populous species of adsorbed oxygen ions are anticipated to be at the testing temperature of 200°C used in this work This is because we will only include in the sensing mechanism equations from Equation8 to Equation13 that are presented in Kröger–Vink notation.
3.7.2. Initial Tests on Adsorption Kinetics with Oxygen (O2) and Nitrogen (N2)
We investigated absorption-dissociation of O
2 on the
Engineered Porosity ZnO surface. The O
2 source was atmospheric air.
Figure 8 illustrates how the electrical resistance (R/R
0) of
Engineered Porosity ZnO changes upon exposure to O
2 and N
2 successively. R
0 denotes electrical resistance as ZnO sensor is maintained under constant O
2 flow, while R denotes the electrical resistance under N
2 flow. N
2 molecules kick off oxygen ions (
from their positions on the surface as soon as O
2 flow is eliminated by N
2 introduction. This causes an injection of electrons back into the ZnO conduction band, which causes ~55% decrease in the electrical resistance (R). Following the removal of the N
2 flow, as shown in
Figure 8 at the 40th minute, and the reintroduction of O
2; the O
2 adsorption-dissociation described in Equations 2 to 7 caused the
Engineered Porosity ZnO electrical resistance to rise to the initial level.
Engineered Porosity ZnO has a great deal of oxygen vacancies and a high degree of network of open porosity.
Figure 8 illustrates how a single exposure to atmospheric air containing 21% O
2 increased electrical resistance by nearly 100%. In other words, large number of oxygen vacancies and high-level connected porosity facilitated large amount of O
2 adsorption and dissociation through the reactions given in Eqs. 2 through 7.
Figure 8.
Adsorption and dissociation of O2 molecules on ZnO and their effect on the electrical resistance at 200°C in air. High purity N2 introduced between Air pulses containing 21% O2.
Figure 8.
Adsorption and dissociation of O2 molecules on ZnO and their effect on the electrical resistance at 200°C in air. High purity N2 introduced between Air pulses containing 21% O2.
3.7.3. Hydrogen Sulfide (H2S) Testing
Figure 9 shows
Engineered Porosity ZnO sensor’s H
2S response at 200°C as the concentration of H
2S varies from a minimum of 0.2 ppm to a maximum of 2.25 ppm. The sensor showed typical n-type semiconductor performance by displaying a decrease in electrical resistance upon exposure to the H
2S reducing gas. During H
2S exposures, chemisorbed oxygen ions interacted with H
2S, releasing electrons back to the conduction band, thus decreasing the electrical resistance according to the mechanisms governed by Equation8 and Equation 9.
Our sensor testing temperature is 200°C, thus we consider
to be the most active adsorbed oxygen specie because
ion formation is favored at ~150–200°C either by direct reduction from
O2 (Equation 3) and/or transformation from
to
(Equation 4 and Equation 5). Equation 8 presents the reaction between H
2S and
. As was previously mentioned, surface oxygen vacancy (
) defect centers associated with ZnO
x helped to facilitate the entire process presented in Equation 8. The reaction between H
2S and adsorbed oxygen (
) produces by product of SO
2 [
17,
18,
46] [
81]. Because SO
2 itself is also reducing agent; SO
2 will interact with the adsorbed oxygen species in the manner outlined in Equation 9.
Figure 9.
Engineered Porosity ZnO tested for sub-ppm H2S at 200°C.
Figure 9.
Engineered Porosity ZnO tested for sub-ppm H2S at 200°C.
At the end of the H2S testing, Engineered Porosity ZnO sensor reintroduced into an oxygen-rich atmosphere relocating the consumed adsorbed oxygen species in accordance with the Equation 2 to Equation 6, as a result we observed the electrical resistance recovered back to its initial value before H2S introduction. SO2 is a by-product of complete combusting of H2S. According to Equation8, H2S reaction with adsorbed oxygen O- produces three free electrons (e-) per reaction, whereas SO2 reaction yields one free electron (e-). The sensing sequence for H2S ends with sulfur trioxide (SO3) as SO2 oxidized to SO3 as seen in Equation 9.
In summary,
Engineered Porosity ZnO abundant with oxygen vacancies showed an n-type sensing behavior for H
2S. ZnO sensor demonstrated exceptional sensing performance, exhibiting high response to 1 ppm H
2S with 45% decrease in electrical resistance yielding a sensor response of 1.7.
Table 2 provides the literature findings for H
2S with the various material-compositions-dopants-precious metals additions etc. along with the results we obtained in our current work.
Examining
Table 2, it is evident that
Engineered Porosity ZnO has distinguished itself with exceptional properties and outperformed the majority of the tabulated literature thanks to its high sensor response and straightforward, simple, and affordable production method. Given that the most recent research in the literature, as shown in
Table 2, makes use of costly precious metal addition, surface alloying, multi-step doping, and/or laborious nano-composites. ZnO sensor can differentiate between extremely low concentrations of H
2S, including 2.25, 1, 0.45, and 0.2 ppm with high accuracy. The ZnO sensor when exposed to CO
2 did not show any sensor response; showed very less toward CO, less than 1% change in electrical resistance.
3.7.4. Nitrogen Oxide (NO2) Testing
Figure 10 displays the testing results for NO
2 at 200°C which also includes the concentrations profiles of O
2 and NO
2 during sensor testing. Combustion processes produce NO
2, which, in fact, coexist in the same environment as another oxidation gas, namely oxygen (O
2). Because of this demanding condition,
Engineered Porosity ZnO was tested for NO
2 while being interfered with by oxygen (O
2). Even with high O
2 interference, ZnO demonstrated exceptional sensing performance at different NO
2 concentrations.
Table 2 compares the NO
2 sensor results from our current work with those from other studies that used different material-compositions-dopants-precious metal additions and/or burdensome nano-composite formations etc. with our results obtained in the current work.
NO
2 introduction increased the electrical resistance due to consumption of conduction band electron as noted from Equation10 to Equation14.
Figure 10 illustrates how the sensor’s electrical resistance increased as the NO
2 concentration was gradually raised, eventually reaches its saturation peak value for 15 ppm of NO
2. As NO
2 concentration decreased, the
Engineered Porosity ZnO electrical resistance decreased to its initial value, emphasizing excellent reversibility and the sensor’s reliability and repeatability.
The
Engineered Porosity ZnO sensor showed maximum response of 15 to 15 ppm NO
2. Even under the interference of 18 ppm of O
2, the sensor managed to detect 3 ppm of NO
2. In comparison to the previous research provided in
Table 2, the exceptional sensor response from
Engineered Porosity ZnO is attributed to a straightforward development process; a high porosity, a well-established tortuosity in microstructure, and a large number of NO
2 pull centers, specifically oxygen vacancies (
). The gas sensing tests of the
Engineered Porosity ZnO sensor with CO
2, CO, toluene and H
2 did not lead to detectable change in the electrical resistance so the results are not presented here.
The reactions for NO
2 sensing are given from Equation 10 to Equation13.
Engineered Porosity ZnO has capability to sense NO
2 even under the presence of O
2 as demonstrated by Eqs. 11 to 13. NO
2 molecules can be adsorbed on oxidized sites on ZnO
x surface (Equation10 to and Equation11), on the other hand; the dissociation and adsorption of NO
2 happens to be at oxygen vacancy sites
() as seen in Eqs. 12 and 13 [
38]. Dissociation and adsorption of NO
2 enabled by the surface oxygen vacancies abundantly found in
Engineered Porosity ZnO. NO
2 dissociates to NO by oxidizing surface oxygen vacancies via donation of
and
ions as given in Eqs. 12 and 13.
As an interference gas, different concentrations of oxygen (O2) were added. This was completed in order to determine whether changes in oxygen gas concentrations could have an impact on the gas sensing mechanism in situations like those found in typical combustion environments with high and fluctuating oxygen content. Because oxygen (O2) interference cannot credibly impact the sensor signal, as demonstrated by the experiments, the Engineered Porosity ZnO is extraordinarily effective at detecting NO2 in realistic environmental settings.
Figure 10.
Engineered Porosity ZnO tested for NO2 at 200°C under O2 interference.
Figure 10.
Engineered Porosity ZnO tested for NO2 at 200°C under O2 interference.
The Table 2 reviews H
2S and NO
2 sensing performances of different compositional and microstructural formations of ZnO including nano-fiber; nano-composites; nano-rods and nano-spheres in various architectures to improve the sensor response. CuO; SnO
2; WS
2 some of the materials were used in addition to the costly precious metals such as platinum (Pt); gold (Au) and palladium (Pd) as surface decoration; doping; additives and/or mechanical alloying. In addition to being difficult; costly; and time-consuming to manufacture; these compositional and microstructural formations have limited applicability for large-scale practice within an acceptable standardization for real-world industrial use.
Engineered Porosity ZnO enriched with surface oxygen vacancies
() showed sensor response of 15 towards 15 ppm of NO
2 and 2.2 to 2.25 ppm of H
2S 200°C.
Engineered Porosity ZnO produced very successful results for NO
2 and H
2S when compared to the results from complex-doped-alloyed-composite material sets presented in
Table 2.
Engineered Porosity ZnO did not have any sensing promoter agent commonly used in the other sensing materials presented in the
Table 2. This work uses engineered porosity designed surface oxygen vacancy (
) enriched ZnO, which
outperforms other sensor platforms-configurations (see
Table 2) in terms of
sensor response magnitude, simplicity, and design robustness. In our work, we reached to such an extraordinary sensor response for pure ZnO material thanks to the heat treatment, engineered porosity-tortuosity network, and enriched surface oxygen vacancy concentration.
Table 2.
Gas Sensing results obtained in this work for Engineered Porosity ZnO, together with literature findings for H2S and NO2 under different gas concentrations and testing temperature. Variety of dopants, additives of other oxide composites and/or precious metals modified ZnO literature findings tabulated for the comparison purposes.
Table 2.
Gas Sensing results obtained in this work for Engineered Porosity ZnO, together with literature findings for H2S and NO2 under different gas concentrations and testing temperature. Variety of dopants, additives of other oxide composites and/or precious metals modified ZnO literature findings tabulated for the comparison purposes.
Sensing Material Composition and Physical State & Form |
Sensor Testing Temperature (°C) |
Test Gas Concentration |
Response Magnitude (Rair/Rgas) for H2S (Rgas/Rair) for NO2
|
Engineered Porosity ZnO (this work) |
200 |
2.25 ppm H2S |
2.2 (This Work) |
CuO/SnO2-ZnO core shell NWs [89] |
25 |
10 ppm H2S |
1.6 |
ZnO/SnO2 nano-fibers [90] |
250 |
50 ppm H2S |
63.3 |
ZnO Nano-rods [91] |
250 |
10 ppm H2S |
20 |
Cu–ZnO nanograins [92] |
250 |
15 ppm H2S |
0.9 |
Au modified ZnO nanowires [93] |
25 |
5 ppm H2S |
6.1 |
Pd–SnO2–ZnO [94] |
25 |
20 ppm H2S |
0.06 |
ZnO Nanowires [95] |
25 |
1 ppm H2S |
1.5 |
Engineered Porosity ZnO (this work) |
200 |
15 ppm NO2 |
15 (This Work) |
ZnO thin film [96] |
250 |
1 ppm NO2
|
2.4 |
ZnO/SnO2-rGO nano-composite [97] |
30 |
5 ppm NO2
|
1.4 |
UV Irradiated Au doped ZnO [40] |
25 |
1 ppm NO2
|
2 |
Co-doped ZnO nano-capsules [98] |
280 |
100 ppm NO2
|
3.86 |
Ni doped ZnO [99] |
200 |
100 ppm NO2
|
2 |
Au decorated ZnO-PANI [100] |
300 |
50 ppm NO2
|
14 |
4. Conclusions
We created a microsensor architecture for the low temperature (200°C) detection of environmentally hazardous gases (NO2 and H2S) in the air using zinc oxide (ZnO) enriched with oxygen vacancy () population and a high degree of interconnected porosity and tortuosity. We established heat treatment procedure leading controlled nano-grain coarsening in order to produce Engineered Porosity ZnO with a network of porosity and tortuosity throughout the active layer thickness.
The sensor demonstrated sensor response against 2.25 ppm of H
2S with a response of 2.2; while for 15 ppm of NO
2; the response was 15 at 200°C.
Engineered Porosity ZnO showed an exceptionally high response at low temperature with an excellent stability-reversibility-repeatability through quick response-recovery times in comparison to the literature findings tabulated in
Table 2. This high sensor response attributed to the two primary reasons; firstly, abundance of oxygen vacancies (
) created by heat treatment procedure and secondly, novel sintered/coarsened grain formation that activated network of porosity and island like connected structures.
Surface analysis of Engineered Porosity ZnO revealed that lattice oxygen attached to ZnO is 22 at.%, whereas 30 at.% of oxygen is found in chemisorbed oxygen. Lattice oxygen in the oxygen vacancy rich phase ZnOx, on the other hand, is 43 at.%. This indicates that the oxygen vacancy- abundant ZnOx controls the majority of the ZnO sensor surface interactions. The O/Zn was found to be 0.73 at.%. We determined x in ZnOx, x=0.73, so the oxygen vacancy abundant phase is ZnO0.73.
Engineered Porosity ZnO detected sub-ppm NO
2 even under oxygen (O
2) interference affect simulating actual sensor operating conditions. The application we accomplished is feasible for industrial level gas sensor production thanks to the straightforward creation of an Engineered Porosity ZnO sensing layer on the sensor chip illustrated in
Figure 1.
In conclusion, heat treatment-coarsening-sintering strategy for creation of oxygen vacancies designed to simultaneously form connected open porous microstructure through the ZnO film led to an exceptional gas sensing properties for NO2 and H2S at 200°C. Surface oxygen vacancies assisted in the capture of target gases and oxygen ions, which propel the sensing to the extraordinary high levels because of the partially empty d-bands achieved in ZnOx.