Next Article in Journal
The Aboveground Biomass Estimation of the Grain for Green Program Stands Using UAV-LiDAR and Sentinel-2 Data
Previous Article in Journal
Optimization of OPM-MEG Layouts with a Limited Number of Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of ZnO Thin Films Doped with Na at Different Percentages for Sensing CO2 in Small Quantities at Room Temperature

by
Marina Stramarkou
1,2,*,
Achilleas Bardakas
2,
Magdalini Krokida
1 and
Christos Tsamis
2
1
School of Chemical Engineering, National Technical University of Athens (NTUA), 15772 Athens, Greece
2
Institute of Nanoscience and Nanotechnology, National Centre for Scientific Research (NCSR) “Demokritos”, 15341 Athens, Greece
*
Author to whom correspondence should be addressed.
Sensors 2025, 25(9), 2705; https://doi.org/10.3390/s25092705
Submission received: 20 March 2025 / Revised: 14 April 2025 / Accepted: 22 April 2025 / Published: 24 April 2025
(This article belongs to the Section Chemical Sensors)

Abstract

:
The objective of this study is the fabrication of sensors which can detect modifications in CO2 concentrations at room temperature, thus indicating the quality or microbial spoilage of food products when incorporated into food packaging. ZnO nanostructures are known for their ability to detect organic gases; however, their effectiveness is limited to high temperatures (greater than 200 °C). To overcome this limitation, sodium (Na) doping is investigated as a way to enhance the sensing properties of ZnO films and lower the working temperature. In this study, undoped and Na-doped ZnO thin films were developed via the sol-gel method with different Na percentages (2.5, 5 and 7.5%) and were deposited via spin coating. The crystal structure, the morphology, and the surface topography of the developed films were characterized by X-ray Diffraction (XRD), Scanning Electron Microscopy (SEM), and Atomic Force Microscopy (AFM), respectively. Furthermore, the response to CO2 was measured by varying its concentration up to 500 ppm at room temperature. All the developed films presented the characteristic diffraction peaks of the ZnO wurtzite hexagonal crystal structure. SEM revealed that the films consisted of densely packed grains, with an average particle size of 58 nm. Na doping increased the film thickness but reduced the surface roughness. Finally, the developed sensors demonstrated very good CO2 sensing properties, with the 2.5% Na-doped sensor having an enhanced sensing performance concerning sensitivity, response, and recovery times. This leads to the conclusion that Na-doped ZnO sensors could be used for the detection of microbial spoilage in food products at room temperature, making them suitable for smart food packaging applications.

1. Introduction

Carbon dioxide (CO2) is a microbial specific marker, as it is a product of the metabolic activities of microbes contaminating food products after unsuitable manipulation and storage and causing food deterioration and, potentially, disease [1]. Specifically, the activity of many microbes that degrade and spoil fish, meat, grains [2], fruits, vegetables, and their juices [3] leads to the formation of CO2, which can change their characteristics and have harmful effects [4]. In addition, CO2 is also a preservative at high concentrations in modified atmosphere food packaging. Therefore, it can indicate not only food spoilage but also the integrity of the packaging. Thus, sensors that detect its presence or changes in its concentration can indicate the quality, freshness, and shelf life of food products [5,6].
In recent years, a variety of CO2 sensors have been developed based on different detection principles, such as optical absorption, electrical resistance, amperometry, etc. However, the practical use of conventional CO2 sensors in food is still limited, as they exhibit several disadvantages, such as high cost, large weight and size, and low durability [7,8,9]. To enable the widespread adoption of CO2 sensors, economical devices that offer simplicity, accuracy, and reliability are needed. To meet these needs, there is an increasing interest in nanomaterial-based sensors due to their cost-effectiveness, robustness to harsh conditions, and stability. Nanomaterials are commonly used for the fabrication of CO2 sensors thanks to their high surface area, which facilitates the adsorption of gas molecules and improves their sensitivity [10,11].
Based on the detection mechanism, CO2 nano-sensors can be categorized into electrochemical, such as chemi-resisitive, capacitive and inductive sensors, and optical, such as non-dispersive infrared, colorimetric, and refractometric sensors [8]. Sensors using electrical transducers are convenient for food applications, as they are fast, economical, and promise easy integration into food packaging materials [6,9].
Electrochemical sensors able to monitor the modification, and more specifically, the increase in the concentration of CO2 can specify the freshness and the quality of food products [5]. With a view to extending the shelf life of fresh and packaged food products, it is beneficial to detect the concentration of CO2 using chemical sensors [12].
Out of the large pool of nanomaterials that could be used for the development of CO2 chemical sensors, zinc oxide (ZnO) is commonly examined [13,14]. ZnO is a II–VI group semiconductor with unique characteristics, such as fast response, low detection limit, high selectivity, and reliable performance. Those properties enable its efficient utilization in various applications, such as chemical sensors for the detection of gases [15].
Although many reports have indicated that ZnO nanostructures can be used for the detection of organic gases, most of them only worked at temperatures greater than 200 °C [11,16]. CO2 is an inert gas with limited oxidizing and reducing properties, which results in poor sensing performance when using semiconductor sensors [17]. This fact critically limits their application in food packaging, where detection temperatures that are as low as possible are required, with ambient temperature being ideal.
Doping ZnO with suitable materials, such as metal atoms and metal oxides like gold (Au) [18], cadmium (Cd) [12], lanthanum (La) [19], aluminum (Al) [20], calcium (Ca) [21], etc. is a method to enhance the sensing behavior of ZnO thin films, lowering the working temperature and at the same time achieving high reliability [20,22,23]. Via the suitable doping of ZnO with metals, its properties can be changed and regulated [24]. The replacement of Zn2+ ions can occur through either lower (group 1) or higher (group 3) valence ion dopants, which increase carrier concentration and lower resistivity (p- or n-type doping). Alternatively, isovalent dopants (i.e., Ca2+) with larger ionic radii can be added to distort the lattice and improve CO2 adsorption [9,25].
Sodium (Na) is a suitable group 1 acceptor with a relatively high hole concentration and shallow substitutional level. Due to its shallow acceptor production, Na is an excellent substitute for Zn as a p-type dopant for ZnO and has gained interest, making it a promising candidate for the fabrication of various p-n junction devices [23,26]. Na also reduces oxygen vacancy density, which is critical in achieving stable p-type conduction [27].
To date, various types of CO2 chemi-resistive sensors based on ZnO thin film structures have been fabricated employing a range of fabrication methods, adjusting different process parameters, and doping with one or more materials in order to enhance the CO2 sensing features. However, as shown in Table 1, most of these developed sensors are limited in one of two ways: they can either detect CO2 only at high concentrations (>1000 ppm) [21,28,29,30] or they require elevated temperatures (>100 °C) to detect lower concentrations [12,31,32,33].
The objective of this study is the fabrication of undoped ZnO and 2.5, 5, and 7.5% Na-doped ZnO thin films, which can sense CO2 gas in low concentrations at room temperature (25 °C), thus being able to be incorporated into food packaging and to inform about the quality of food products. The crystal structure of the developed films was characterized by X-ray Diffraction (XRD), their morphology and thickness were investigated through Field Emission Scanning Electron Microscopy (FE-SEM), and their surface topography was evaluated using Atomic Force Microscopy (AFM). The response to CO2 was measured by modulating its concentration up to 500 ppm at room temperature using an appropriate experimental setup. To the best of our knowledge, this is the first work published that studied ZnO sensors which are able to detect CO2 at very low concentrations, at room temperature, and without the use of UV light, for food packaging applications.

2. Materials and Methods

2.1. Sol-Gel Preparation, Film Deposition, and Substrate Development

The preparation of the sol-gels, as well as the deposition of the films, were based on the study of Basyooni et al. (2017) and further expanded regarding the concentration envelope of the dopant, including the percentages of 5 and 7.5% Na, using a SiO2 substrate [23].
The preparation of the undoped ZnO sol-gels involved the dissolution of zinc acetate dihydrate (source material) (Sigma-Aldrich, Darmstadt, Germany) in 2-methoxyethanol (2ME) solvent (Sigma-Aldrich, Darmstadt, Germany) in order to obtain a 0.5 M solution, with the addition of 0.5 M ethanolamine (MEA) (Sigma-Aldrich, Darmstadt, Germany) as a stabilizing agent. For the preparation of the Na-doped sol-gels, sodium acetate (dopant material) (Sigma-Aldrich, Darmstadt, Germany) was added to the above solutions, reaching a final concentration of 2.5, 5 and 7.5 at % Na. The solutions were magnetically stirred for 120 min at 60 °C and were left at room-temperature overnight in a dark environment.

2.2. Sensor Fabrication

The process started with the cleaning of a four-inch n-type Si (100) wafer using a piranha solution (H2O2/H2SO4 (VSLI-grade, Microchemicals GmbH, Ulm, Germany), 1:1 vol.) (Figure 1, step 1) in order to remove possible organic contamination, followed by rinsing with deionized (DI) water and blow drying with nitrogen (N2). After cleaning, a 100 nm-thick silicon dioxide (SiO2) was grown on the wafer using dry thermal oxidation, followed by another cleaning step before further processing (Figure 1, step 2). Subsequently, negative lithography was carried out by spin coating a 1.3 μm thick AZ-5214E (Microchemicals GmbH, Ulm, Germany) photo-resistant thin film on the SiO2 at 5000 rpm, followed by a pre-bake at 110 °C for 90 s. Exposure under UV light was performed using a SUSS Microtec MA6 mask aligner for 500 s (Step 3), using a quartz photomask containing the device features. After exposure, a post-exposure bake was performed at 120 °C for 90 s. Following that, a flood exposure was performed for 3 min in order to perform the image reversal process. The development of the exposed pattern was performed by soaking the wafer in the AZ-726 MIF (VLSI-grade, Microchemicals GmbH, Ulm, Germany) developer for 60 s. Afterward, direct current (DC) magnetron sputtering was employed in order to deposit a 10 nm thick titanium (Ti) adhesion layer and a 50 nm thick gold (Au) layer (Figure 1, step 4). Following the deposition, the lift-off process (Figure 1, step 5) was performed in an acetone (EMSURE® ACS, ISO, Reag. Ph Eur, Merck KGaA, Darmstadt, Germany) bath for 15 min without ultrasonication, defining the final interdigitated electrode (IDE) geometry, followed by rinsing with isopropanol (EMSURE® ACS, ISO, Reag. Ph Eur, Merck KGaA, Darmstadt, Germany) and DI water and blow drying with N2. Each sample was cleaved from the processed wafer using a diamond scribe containing three sensors. The final fabrication step was the deposition of the ZnO thin films using spin coating. Prior to the deposition, the area around the IDEs was masked using Kapton® tape in order to deposit the ZnO thin film only at the electrode area. The film deposition was carried out with seven spin coatings (2000 rpm, 30 s) (Figure 1, step 6), followed by thermal treatments (after each coating) using a hotplate (180 °C, 20 min) and a final annealing step at 500 °C for 120 min (Figure 1, step 7). The steps of the whole fabrication process are shown schematically in Figure 1.

2.3. Morphological and Crystallograpic Characterisation

X-ray diffraction (XRD) analysis was used to investigate the microstructure of the doped and undoped ZnO thin film (D8 Advance Bruker, Billerica, MA, USA) using a Cu-Kα (λ = 1.5405 nm) radiation source at a 0.029/s scanning rate in the 2θ range of 20–80 degrees. The crystallite size of the films was calculated using Debye Scherrer’s formula:
D = 0.94 λ β   cos θ
where λ is the wavelength of the X-ray source, β is the full width at half maximum (FWHM) in radians, and θ is the diffraction angle of the highest peak in radians.
The morphological properties and thickness of the films were studied using field emission-scanning electron microscopy (FE-SEM, JEOL JSM-7401F, Tokyo, Japan) operating at a voltage of 2.5 kV in top-view and cross-sectional perspective. The analysis of FE-SEM micrographs was performed with the open-source image analysis tool ImageJ (v. 1.54g).
Surface roughness measurements were performed by atomic force microscopy (AFM) (NTEGRA Prima atomic force microscope, (Spectrum Instruments Ltd., Limerick, Ireland)) in intermittent contact mode. For this purpose, NT-MDT NSGO1 silicon, N-type, antimony doped cantilevers with Au reflective coating and a nominal force constant of 5.1 N/m were used. AFM images were created and analyzed with the software Gwyddion 2.60 (Free and Open-Source software, Department of Nanometrology, Czech Metrology Institute).
A statistical analysis was performed with one-way and factorial analysis of variance (ANOVA) in order to analyze the differences between the four developed sensors concerning their particle size and thickness using STATISTICA software (v.13.6 StatSoft®Inc., Palo Alto, CA, USA).

2.4. Electrical Characterization

The sensing performance of the developed sensors was evaluated at different CO2 concentrations: 50, 125, 250, 330, and 500 ppm. After the fabrication process, the sample containing the three CO2 sensors was mounted on a printed circuit board (PCB) and connected via wire bonding. During the measurements, the sensors were connected to a Keithley 2400 (Tektronix, Berkshire, UK) source measure unit (SMU) and placed in a sealed Teflon® (Manchester, UK) chamber at room temperature (25 °C). The control of the CO2 content was done by providing a mixture of synthesized dry air (N2 80%, O2 20%) and N2 gas with 0.1% CO2 through Brooks® 5800-S mass flow controllers (MFC) (Brooks Instrument, Hatfield, PA, USA). A bias of 2 V was constantly supplied by the SMU, and the current of the sensor was recorded. The response (%) of the sensors was calculated as:
S % = R a R g R a ×   100
where S is response, Ra is the resistance in the absence of CO2, and Rg is the resistance in the presence of different concentrations of CO2. The response time was equal to the time the sensors needed to achieve 90% of the total signal change from the moment they were exposed to a specific CO2 concentration. The recovery time was equal to the time required for the sensors’ signal to return to 90% of their initial value after the CO2 concentration value was zero [47].

3. Results

3.1. X-Ray Diffraction (XRD)

The X-ray diffraction patterns of undoped and Na-doped ZnO thin films are compared in Figure 2.
The detected diffraction peaks, which were attributed to the (100), (002), (101), (102), (110), (103), (200), (112), and (201) facets, are consistent with the ZnO wurtzite hexagonal crystal structure (JCPDS 36-1451). The high and wide peak that arose at 2θ~69.56 was attributed to the Si substrate, probably due to the small thickness of the sensing films, which permitted the X-ray to penetrate the sensing layer and reach the substrate. All the samples showed the preferred orientations along the facets (100), (002), and (101). The doping with Na did not result in extra diffraction peaks except for those of pure ZnO, but clearly intensified the c-axis (002) and decreased the intensity of (100) and (101) facets, especially at lower Na concentrations (2.5%). This was also observed in ZnO thin films doped with calcium [21], aluminum, and magnesium [30]. The incorporation of Na led to a slight shift of the diffraction peaks to higher 2θ values (from 34.63 to 34.65) because of the larger ionic radius of Na (0.95 nm) compared to Zn (0.74 nm), thus forming lattice defects [21]. In addition, the average crystallite size (d) of ZnO thin films decreased from 39 nm to 35 nm and 34 nm with the incorporation of 2.5% and 5% Na, respectively. However, the further increase in dopant concentration (7.5% Na) increased the crystallite size to 38 nm. The non-linear change of crystallite size depending on the percentage of the dopant was also reported in a study by Altun et al. (2021) [12].

3.2. Field Emission-Scanning Electron Microscopy (FE-SEM)

The results of the structural analysis performed by FE-SEM are presented in Table 2. As observed, both ZnO and Na-doped ZnO thin films consisted of dense grains from agglomerated nanoparticles. Doping with Na did not significantly change the morphology or the mean particle size, which were statistically similar. Regarding the size distribution, 2.5% Na-doped films exhibited a wider range (24.4–114.7 nm), with most of the particles, however, being well concentrated in the 47–58.3 nm region, in contrast to the pure ZnO films, which had a slightly narrower spectrum but the majority of the grains were in a larger range (35.8–73.4 nm). This phenomenon was also observed in previous research by Benzitouni et al. (2017) and was attributed to the formation of grain aggregates due to the higher mobility caused by doping [48]. The growth process of large particles is known as Ostwald ripening: large particles grow at the expense of small particles [49]. The ripening process was even more evident in films with a higher doping percentage (7.5% Na), where several possibly aggregated grains with a size of 101.6–112.2 nm appeared. Finally, it was observed that after 2.5% Na doping, the film thickness increased by about 30 nm. As the doping percentage increased, the films became thicker, with the 7.5% Na-doped ZnO films showing the highest thickness.

3.3. Atomic Force Microscopy (AFM)

The two- and three-dimensional topography images, as well as the root mean square (RMS) roughness for the sensor films developed in this research, are shown in Table 3.
In general, all the sensing films showed the characteristic surface of ZnO, which was composed of spherical particles at various scales. Comparing the topographies of the plain ZnO film and the Na-doped films, the latter exhibited a smoother surface, which was confirmed by the reduction of the RMS surface roughness. The creation of a smoother structure and surface after doping was also highlighted in the research of Benzitouni (2017) and Banerjee (2010), where cobalt and aluminum were incorporated into the ZnO film, respectively [48,50]. In fact, in both studies, it was concluded that the roughness depended, to a significant extent, on the percentage of the dopant, and that while the doping with a small percentage led to a significant reduction of the RMS roughness, the incorporation of larger percentages led to an increase of its value, as was the case in the present study.
The combination of the results of the two microscopy methods, FE-SEM and AFM, led to the conclusion that the surface uniformity was strongly influenced by the thickness of the thin film and the arrangement of the grain particles. Increasing the film thickness and its particle size led to a better surface uniformity and a decrease in the roughness value [51]. Xu (2012) reported similar results on the relationship between ZnO thin film thickness and RMS surface roughness values [52].

3.4. CO2 Measurements

The sensing response of the developed sensors is shown in Table 4. The response of the sensors was evaluated for various CO2 values (50, 125, 250, 330 and 500 ppm) under a constant voltage of 2 V and at room temperature (25 °C). During the experimental procedure, a constant voltage was applied to the sensor terminals while the sensor current value and time were recorded. The measurement began at 0% CO2 concentrations within the chamber; then with an appropriate combination of flow through the two controllers, the CO2 concentrations were increased. After each CO2 step, a zero concentration step followed so that the sensor response and recovery time could be subsequently calculated. The net response (net signal) of the four developed sensors was determined by establishing a baseline at 0% CO2 concentration and subtracting it from the recorded signals at different CO2 levels.
The response of the sensors, as well as the response and recovery times as a function of CO2 concentration, are shown in Figure 3.
In general, the main mechanism of CO2 gas detection by ZnO is based on the adsorption of CO2 molecules on the ZnO surface, resulting in a change in the electrical resistance of the surface [53].
More specifically, in the presence of atmospheric air, oxygen molecules (O2) are adsorbed on the ZnO surface, transferring electrons (e) from its conduction band and forming oxygen ions O2, O, and O2−, depending on the temperature (low, intermediate, or high). This results in the electron concentration decreasing and an electron depletion zone being created on the surface of each ZnO molecule [28]. The following equations describe the whole process as a function of temperature [54]:
O2 (gas) → O2 (adsorbed)
O2 (adsorbed) + e(surface) ↔ O2 (adsorbed) (<100 °C)
O2 (adsorbed) + 2e(surface) ↔ 2O (adsorbed) (100–300 °C)
O (adsorbed) + e(surface) ↔ O2− (adsorbed) (>300 °C)
When the ZnO surface is exposed to CO2 gas, the adsorbed anions O2, O, and O2− react with the CO2 molecules, which extract electrons from the conduction band of ZnO and form carbonate compounds, such as (CO3)2− [55]. Consequently, chemical bonding and charge transfer occur between the metal oxide surface and the adsorbed gas. Since ZnO is an n-type semiconductor, charge transfer reduces the concentration of electrons on the ZnO surface, leading to an increase in the electrical resistance of the sensor, as shown in Table 4 [56]. The second phase of the process occurs according to the following Equations [57]:
CO2 (gas) + O (adsorbed) → (CO3)2−(adsorbed)
(CO3)2−(adsorbed) → CO2 (gas) + ½ O2(gas)
As seen in Table 4, the response of the sensors increased by increasing the CO2 content. The exact dependence of the sensor response on the different CO2 levels can be found in the equation shown in each figure of the response/CO2 concentration. In general, it was noticed that as the CO2 concentration increased, the response reached a relative plateau, especially in the cases of the sensors with 2.5 and 5% Na.
Comparing the performance of the plain ZnO sensor and the sensors with 2.5% and 5% Na, it was observed that the doping with minor concentrations of dopant contributed to an obvious improvement in the sensitivity and in the response and recovery times. Especially the sensor with 2.5% reduces the signal noise at zero CO2 concentration, whereas the sensor with 5% Na shortens the response and recovery times the most, which is a favorable effect for sensors. These improvements are due to the fact that by adding Na+ ions the microstructure of the pure ZnO changes and the conductivity increases. ZnO has a narrow hexagonal lattice, which is “open” to some extent since Zn atoms occupy half of the tetrahedral sites and all the octahedral sites are empty. Therefore, ZnO offers many sites to “accommodate” intrinsic defects and extrinsic impurities. When Na+ ions replace zinc Zn2+ sites, oxygen vacancies (VO•) are created. CO2 is then adsorbed through the formation of Na2CO3. The above-mentioned changes are described by the following Kröger Vink Equations:
Na+ ion substitution for Zn2+ in the lattice:
N a N a Z N + V
CO2 adsorption and formation of Na2CO3:
N a + C O 2 a b s o r b e d N a 2 C O 3
CO2 adsorption and interaction with oxygen vacancies:
C O 2 + V O C O 2 + V O x
Ionization of oxygen vacancies:
V O V O + e
where, O: oxygen, V: vacancy, : positive charge and e′: negative charge.
Therefore, replacing Zn2+ with Na+ in the ZnO crystal lattice can create additional oxygen vacancies V O and defects, with Na-ZnO containing heavier concentrations of oxygen vacancies than ZnO. Subsequently, these additional vacancies increase electron concentration and CO2 adsorption, which enhances the response of the doped sensor and enables satisfactory detection of CO2 at room temperature [23]. The enhancement of sensing performance by 2.5% Na-doped ZnO sensor can be also attributed to the smaller size of the crystallites estimated via the XRD analysis results, which was also observed in the work of Petrov et al. (2021) [58].
Furthermore, the doping with a higher percentage of Na causes difficulty in detecting low amounts of CO2 since the sensor with 7.5% Na was not able to detect the amounts of 50 and 125 ppm of CO2. This is due to several factors, which are confirmed by FE-SEM and AFM morphology. Doping with 5 and 7.5% Na leads to a minimal increase in particle size due to agglomerations and a significant increase in film thickness, thus reducing both the CO2 adsorption sites and the surface-to-volume ratio. Therefore, the CO2 gas adsorption on the ZnO film is impeded [34]. The deterioration of the sensor response when incorporating higher percentages of dopants in ZnO has been observed in various studies [12,31,32,59,60].

4. Conclusions

In conclusion, Na doping at a low concentration of 2.5% in ZnO significantly improved the CO2 sensing performance of ZnO films, making them effective for detecting microbial spoilage in food products at room temperature. While pure ZnO nanostructures can detect CO2 at high temperatures higher than 200 °C, the addition of Na overcame the limitation and had various effects. The 2.5% Na-doped ZnO sensor demonstrated the most favorable properties, including reduced crystallite size, increased film thickness, and a smoother, more uniform surface structure. Finally and most importantly, Na doping at 2.5% and 5% enhanced the sensor performance by allowing the detection of very low concentrations of CO2, even 50 ppm, increasing the sensitivity and reducing the response and recovery times, while eliminating the “noise”. Overall, the developed sensors showed satisfactory CO2 sensing properties, with the 2.5% and 5% Na doped sensors having the best performance. These results confirm that the developed sensors can successfully be integrated into a smart packaging system, providing a real-time and reliable method to monitor freshness in food products at room temperature.

Author Contributions

Conceptualization, C.T.; methodology, M.S. and A.B.; software, M.S. and A.B.; validation, M.S. and C.T.; formal analysis, M.S.; investigation, M.S. and C.T.; resources, M.S., M.K. and C.T.; data curation, M.S.; writing—original draft preparation, M.S.; writing—review and editing, M.S., A.B. and C.T.; visualization, M.S. and A.B.; supervision, M.K. and C.T.; project administration, M.K. and C.T.; funding acquisition, M.K. and C.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Stavros Niarchos Foundation through the Industrial Research Fellowship Program at NCSR “Demokritos” and NTUA.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author (Marina Stramarkou’ mstrmarkou@chemeng.ntua.gr).

Acknowledgments

The authors would like to acknowledge the help and technical support from the Nanotechnology and Microsystems Laboratory of the Institute of Nanoscience and Nanotechnology (INN) at NCSR “Demokritos”.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. de Cássia Martins Salomão, B. Pathogens and Spoilage Microorganisms in Fruit Juice: An Overview. In Fruit Juices; Elsevier Inc.: Amsterdam, The Netherlands, 2018; pp. 291–308. [Google Scholar] [CrossRef]
  2. Mustafa, F.; Andreescu, S. Chemical and Biological Sensors for Food-Quality Monitoring and Smart Packaging. Foods 2018, 7, 168. [Google Scholar] [CrossRef] [PubMed]
  3. Mafe, A.N.; Edo, G.I.; Makia, R.S.; Joshua, O.A.; Akpoghelie, P.O.; Gaaz, T.S.; Jikah, A.N.; Yousif, E.; Isoje, E.F.; Igbuku, U.A.; et al. A review on food spoilage mechanisms, food borne diseases and commercial aspects of food preservation and processing. Food Chem. Adv. 2024, 5, 100852. [Google Scholar] [CrossRef]
  4. Joshi, N.; Pransu, G.; Adam Conte-Junior, C. Critical Review and Recent Advances of 2D Materials-Based Gas Sensors for Food Spoilage Detection. Crit. Rev. Food Sci. Nutr. 2022, 63, 10536–10559. [Google Scholar] [CrossRef]
  5. Gobbi, E.; Falasconi, M.; Concina, I.; Mantero, G.; Bianchi, F.; Mattarozzi, M.; Musci, M.; Sberveglieri, G. Electronic Nose and Alicyclobacillus Spp. Spoilage of Fruit Juices: An Emerging Diagnostic Tool. Food Control 2010, 21, 1374–1382. [Google Scholar] [CrossRef]
  6. Weston, M.; Geng, S.; Chandrawati, R. Food Sensors: Challenges and Opportunities. Adv. Mater. Technol. 2021, 6, 2001242. [Google Scholar] [CrossRef]
  7. Pavase, T.R.; Lin, H.; Shaikh, Q.; Hussain, S.; Li, Z.; Ahmed, I.; Lv, L.; Sun, L.; Shah, S.B.H.; Kalhoro, M.T. Recent Advances of Conjugated Polymer (CP) Nanocomposite-Based Chemical Sensors and Their Applications in Food Spoilage Detection: A Comprehensive Review. Sens. Actuators B Chem. 2018, 273, 1113–1138. [Google Scholar] [CrossRef]
  8. Rezk, M.Y.; Sharma, J.; Gartia, M.R. Nanomaterial-Based CO2 Sensors. Nanomaterials 2020, 10, 2251. [Google Scholar] [CrossRef]
  9. Stramarkou, M.; Bardakas, A.; Krokida, M.; Tsamis, C. ZnO-Based Chemi-Resistive Sensors for CO2 Detection: A Review. Sens. Rev. 2022, 42, 682–706. [Google Scholar] [CrossRef]
  10. Kotbi, A.; Benyoussef, M.; Ressami, E.M.; Lejeune, M.; Lakssir, B.; Jouiad, M. Gas Sensors Based on Exfoliated G-C3N4 for CO2 Detection. Chemosensors 2022, 10, 470. [Google Scholar] [CrossRef]
  11. Dhahri, R.; Leonardi, S.G.; Hjiri, M.; El Mir, L.; Bonavita, A.; Donato, N.; Iannazzo, D.; Neri, G. Enhanced Performance of Novel Calcium/Aluminum Co-Doped Zinc Oxide for CO2 Sensors. Sens. Actuators B Chem. 2017, 239, 36–44. [Google Scholar] [CrossRef]
  12. Altun, B.; Karaduman Er, I.; Çağırtekin, A.O.; Ajjaq, A.; Sarf, F.; Acar, S. Effect of Cd Dopant on Structural, Optical and CO2 Gas Sensing Properties of ZnO Thin Film Sensors Fabricated by Chemical Bath Deposition Method. Appl. Phys. A Mater. Sci. Process 2021, 127, 687. [Google Scholar] [CrossRef]
  13. Franco, M.A.; Conti, P.P.; Andre, R.S.; Correa, D.S. A Review on Chemiresistive ZnO Gas Sensors. Sens. Actuators Rep. 2022, 4, 100100. [Google Scholar] [CrossRef]
  14. Bagheri, F.; Haratizadeh, H.; Ahmadi, M. Improving CO2 Sensing and P-n Conductivity Transition under UV Light by Chemo-Resistive Sensor Based on ZnO Nanoparticles. Ceram. Int. 2024, 50, 1497–1504. [Google Scholar] [CrossRef]
  15. Gong, J.; Li, Y.; Chai, X.; Hu, Z.; Deng, Y. UV-Light-Activated ZnO Fibers for Organic Gas Sensing at Room Temperature. J. Phys. Chem. C 2010, 114, 1293–1298. [Google Scholar] [CrossRef]
  16. Ghobadifard, M.; Khelghati, M.; Zamani, E.; Maleki, Q.; Farhadi, S.; Aslani, A. Zinc Oxide Nano-Crystals Assisted for Carbon Dioxide Gas Sensing; Prepared by Solvothermal and Sonochemical Methods. Iran. Chem. Commun. 2015, 3, 26–40. [Google Scholar]
  17. Duan, X.; Jiang, Y.; Liu, B.; Duan, Z.; Zhang, Y.; Yuan, Z.; Tai, H. Enhancing the Carbon Dioxide Sensing Performance of LaFeO3 by Co Doping. Sens. Actuators B Chem. 2024, 402, 135136. [Google Scholar] [CrossRef]
  18. González-Garnica, M.; Galdámez-Martínez, A.; Malagón, F.; Ramos, C.D.; Santana, G.; Abolhassani, R.; Kumar Panda, P.; Kaushik, A.; Mishra, Y.K.; Karthik, T.V.K.; et al. One Dimensional Au-ZnO Hybrid Nanostructures Based CO2 Detection: Growth Mechanism and Role of the Seed Layer on Sensing Performance. Sens. Actuators B Chem. 2021, 337, 129765. [Google Scholar] [CrossRef]
  19. Abdelkarem, K.; Saad, R.; El Sayed, A.M.; Fathy, M.I.; Shaban, M.; Hamdy, H. Design of High-Sensitivity La-Doped ZnO Sensors for CO2 Gas Detection at Room Temperature. Sci. Rep. 2023, 13, 18398. [Google Scholar] [CrossRef]
  20. Mishra, M.; Kumar, M. Effects of Al Doping on Structural, Optical, Morphological, and Low-Concentration CO2 Gas Sensing Properties of ZnO Nanoparticles Synthesized by Sol-Gel Method. Phys. Scr. 2024, 99, 125929. [Google Scholar] [CrossRef]
  21. Ghosh, A.; Zhang, C.; Zhang, H.; Shi, S. CO2 Sensing Behavior of Calcium-Doped ZnO Thin Film: A Study to Address the Cross-Sensitivity of CO2 in H2 and CO Environment. Langmuir 2019, 35, 10267–10275. [Google Scholar] [CrossRef]
  22. Soltabayev, B.; Raiymbekov, Y.; Nuftolla, A.; Turlybekuly, A.; Yergaliuly, G.; Mentbayeva, A. Sensitivity Enhancement of CO2 Sensors at Room Temperature Based on the CZO Nanorod Architecture. ACS Sens. 2024, 9, 1227–1238. [Google Scholar] [CrossRef] [PubMed]
  23. Basyooni, M.A.; Shaban, M.; El Sayed, A.M. Enhanced Gas Sensing Properties of Spin-Coated Na-Doped ZnO Nanostructured Films. Sci. Rep. 2017, 7, srep41716. [Google Scholar] [CrossRef] [PubMed]
  24. Zhu, L.; Zeng, W. Room-Temperature Gas Sensing of ZnO-Based Gas Sensor: A Review. Sens. Actuators A Phys. 2017, 267, 242–261. [Google Scholar] [CrossRef]
  25. Pantò, F.; Leonardi, S.G.; Fazio, E. CO2 Sensing Properties of Electro-Spun Ca-Doped ZnO Fi Bres. Nanotechnology 2018, 29, 305501. [Google Scholar] [CrossRef]
  26. Elsayed, M.H.; Elmorsi, T.M.; Abuelela, A.M.; Hassan, A.E.; Alhakemy, A.Z.; Bakr, M.F.; Chou, H.H. Direct Sunlight-Active Na-Doped ZnO Photocatalyst for the Mineralization of Organic Pollutants at Different PH Mediums. J. Taiwan Inst. Chem. Eng. 2020, 115, 187–197. [Google Scholar] [CrossRef]
  27. Akcan, D.; Gungor, A.; Arda, L. Structural and Optical Properties of Na-Doped ZnO Films. J. Mol. Struct. 2018, 1161, 299–305. [Google Scholar] [CrossRef]
  28. Dhawale, D.S.; Lokhande, C.D. Chemical Route to Synthesis of Mesoporous ZnO Thin Films and Their Liquefied Petroleum Gas Sensor Performance. J. Alloys Compd. 2011, 509, 10092–10097. [Google Scholar] [CrossRef]
  29. Pohle, R.; Tawil, A.; Mrotzek, C.; Fleischer, M. CO2 Sensing by Work Function Readout of ZnO Based Screen Printed Films. In Proceedings of the 2013 Transducers and Eurosensors XXVII: The 17th International Conference on Solid-State Sensors, Actuators and Microsystems, TRANSDUCERS and EUROSENSORS 2013, Barcelona, Spain, 16–20 June 2013. [Google Scholar] [CrossRef]
  30. Jaballah, S.; Dahman, H.; El Mir, L.; Neri, G.; Donato, N. CO and CO2 Sensing by Al-Mg-ZnO Based Conductometric Sensor. In Proceedings of the IEEE Medical Measurements and Applications, MeMeA 2020—Conference Proceedings, Bari, Italy, 1 June–1 July 2020. [Google Scholar] [CrossRef]
  31. Shokry Hassan, H.; Kashyout, A.B.; Morsi, I.; Nasser, A.A.A.; Ali, I. Synthesis, Characterization and Fabrication of Gas Sensor Devices Using ZnO and ZnO:In Nanomaterials. Beni Suef Univ. J. Basic. Appl. Sci. 2014, 3, 216–221. [Google Scholar] [CrossRef]
  32. Motaung, D.E.; Kortidis, I.; Mhlongo, G.H.; Duvenhage, M.M.; Swart, H.C.; Kiriakidis, G.; Ray, S.S. Correlating the Magnetism and Gas Sensing Properties of Mn-Doped ZnO Films Enhanced by UV Irradiation. RSC Adv. 2016, 6, 26227–26238. [Google Scholar] [CrossRef]
  33. Hunge, Y.M.; Yadav, A.A.; Kulkarni, S.B.; Mathe, V.L. A Multifunctional ZnO Thin Film Based Devises for Photoelectrocatalytic Degradation of Terephthalic Acid and CO2 Gas Sensing Applications. Sens. Actuators B Chem. 2018, 274, 1–9. [Google Scholar] [CrossRef]
  34. Samarasekara, P.; Yapa, N.U.S.; Kumara, N.T.R.N.; Perera, M.V.K. CO2 Gas Sensitivity of Sputtered Zinc Oxide Thin Films. Bull. Mater. Sci. 2007, 30, 113–116. [Google Scholar] [CrossRef]
  35. Lupan, O.; Chow, L.; Shishiyanu, S.; Monaico, E.; Shishiyanu, T.; Şontea, V.; Roldan Cuenya, B.; Naitabdi, A.; Park, S.; Schulte, A. Nanostructured Zinc Oxide Films Synthesized by Successive Chemical Solution Deposition for Gas Sensor Applications. Mater. Res. Bull. 2009, 44, 63–69. [Google Scholar] [CrossRef]
  36. Kannan, P.K.; Saraswathi, R.; Rayappan, J.B.B. CO2 Gas Sensing Properties of DC Reactive Magnetron Sputtered ZnO Thin Film. Ceram. Int. 2014, 40, 13115–13122. [Google Scholar] [CrossRef]
  37. Dhahri, R.; Hjiri, M.; El Mir, L.; Fazio, E.; Neri, F.; Barreca, F.; Donato, N.; Bonavita, A.; Leonardi, S.G.; Neri, G. ZnO:Ca Nanopowders with Enhanced CO2 Sensing Properties. J. Phys. D Appl. Phys. 2015, 48, 255503. [Google Scholar] [CrossRef]
  38. Shukla, R.K.; Srivastava, A.; Kumar, N.; Pandey, A.; Pandey, M. Optical and Sensing Properties of Cu Doped ZnO Nanocrystalline Thin Films. J. Nanotechnol. 2015, 2015, 172864. [Google Scholar] [CrossRef]
  39. Martínez, L.; Holguín-Momaca, J.T.; Karthik, T.V.K.; Olive-Méndez, S.F.; Campos-Alvarez, J.; Agarwal, V. Sputtering Temperature Dependent Growth Kinetics and CO2 Sensing Properties of ZnO Deposited over Porous Silicon. Superlattices Microstruct. 2016, 98, 8–17. [Google Scholar] [CrossRef]
  40. Hjiri, M.; Zahmouli, N.; Dhahri, R.; Leonardi, S.G.; El Mir, L.; Neri, G. Doped-ZnO Nanoparticles for Selective Gas Sensors. J. Mater. Sci. Mater. Electron. 2017, 28, 9667–9674. [Google Scholar] [CrossRef]
  41. Karthik, T.V.K.; Martinez, L.; Agarwal, V. Porous Silicon ZnO/SnO2 Structures for CO2 Detection. J. Alloys Compd. 2018, 731, 853–863. [Google Scholar] [CrossRef]
  42. Ghanbari Shohany, B.; Motevalizadeh, L.; Ebrahimizadeh Abrishami, M. Investigation of ZnO Thin-Film Sensing Properties for CO2 Detection: Effect of Mn Doping. J. Theor. Appl. Phys. 2018, 12, 219–225. [Google Scholar] [CrossRef]
  43. Ghosh, A.; Zhang, C.; Shi, S.; Zhang, H. High Temperature CO2 Sensing and Its Cross-Sensitivity towards H2 and CO Gas Using Calcium Doped ZnO Thin Film Coated Langasite SAW Sensor. Sens. Actuators B Chem. 2019, 301, 126958. [Google Scholar] [CrossRef]
  44. Bhowmick, T.; Ghosh, A.; Nag, S.; Majumder, S.B. Sensitive and Selective CO2 Gas Sensor Based on CuO/ZnO Bilayer Thin-Film Architecture. J. Alloys Compd. 2022, 903, 163871. [Google Scholar] [CrossRef]
  45. Xia, Y.; Pan, A.; Su, Y.-Q.; Zhao, S.; Li, Z.; Davey, A.K.; Zhao, L.; Maboudian, R.; Carraro, C. In-Situ Synthesized N-Doped ZnO for Enhanced CO2 Sensing: Experiments and DFT Calculations. Sens. Actuators B Chem. 2022, 357, 131359. [Google Scholar] [CrossRef]
  46. Mishra, M.; Banga, V.P.; Kumar, M.; Gupta, M. Effect of Aging on Transmittance, and Effect of Annealing Temperature on CO2 Sensing of ZnO Thin Film Deposited by Spin Coating. e-Prime-Adv. Electr. Eng. Electron. Energy 2024, 7, 100405. [Google Scholar] [CrossRef]
  47. Wei, A.; Pan, L.; Huang, W. Recent Progress in the ZnO Nanostructure-Based Sensors. Mater. Sci. Eng. B Solid. State Mater. Adv. Technol. 2011, 176, 1409–1421. [Google Scholar] [CrossRef]
  48. Benzitouni, S.; Zaabat, M.; Ebothe, J.; Boudine, B.; Coste, R. The Use of Advanced Atomic Force Microscopy for the Quantitative Nanomechanical Characterization of Co-Doped ZnO Thin Films. Chin. J. Phys. 2017, 55, 2458–2467. [Google Scholar] [CrossRef]
  49. Behera, D.; Acharya, B.S. Nano-Star Formation in Al-Doped ZnO Thin Film Deposited by Dip-Dry Method and Its Characterization Using Atomic Force Microscopy, Electron Probe Microscopy, Photoluminescence and Laser Raman Spectroscopy. J. Lumin. 2008, 128, 1577–1586. [Google Scholar] [CrossRef]
  50. Banerjee, P.; Lee, W.J.; Bae, K.R.; Lee, S.B.; Rubloff, G.W. Structural, Electrical, and Optical Properties of Atomic Layer Deposition Al-Doped ZnO Films. J. Appl. Phys. 2010, 108, 043504. [Google Scholar] [CrossRef]
  51. Haarindraprasad, R.; Hashim, U.; Gopinath, S.C.B.; Kashif, M.; Veeradasan, P.; Balakrishnan, S.R.; Foo, K.L.; Poopalan, P.; Mishra, Y.K. Low Temperature Annealed Zinc Oxide Nanostructured Thin Film-Based Transducers: Characterization for Sensing Applications. PLoS ONE 2015, 10, e0132755. [Google Scholar] [CrossRef]
  52. Xu, D.; Jiang, B.; Jiao, L.; Cui, F.D.; Xu, H.X.; Yang, Y.T.; Yu, R.H.; Cheng, X.N. Sol-Gel Synthesis of Y2O3-Doped ZnO Thin Films Varistors and Their Electrical Properties. Trans. Nonferrous Met. Soc. China (Engl. Ed.) 2012, 22 (Suppl. S1), s110–s114. [Google Scholar] [CrossRef]
  53. Ramos-Ramón, J.A.; Bogireddy, N.K.R.; Giles Vieyra, J.A.; Karthik, T.V.K.; Agarwal, V. Nitrogen-Doped Carbon Dots Induced Enhancement in CO2 Sensing Response from ZnO–Porous Silicon Hybrid Structure. Front. Chem. 2020, 8, 291. [Google Scholar] [CrossRef]
  54. Mirzaei, A.; Lee, J.H.; Majhi, S.M.; Weber, M.; Bechelany, M.; Kim, H.W.; Kim, S.S. Resistive Gas Sensors Based on Metal-Oxide Nanowires. J. Appl. Phys. 2019, 126, 241102. [Google Scholar] [CrossRef]
  55. Hernandez, A.G.; Kudriavtsev, Y.; Karthik, T.V.K.; Asomoza, R. Optimization of Ge Substrates for ZnO Deposition and Their Application for CO2 Detection. J. Mater. Sci. Mater. Electron. 2019, 30, 6660–6668. [Google Scholar] [CrossRef]
  56. Tit, N.; Othman, W.; Shaheen, A.; Ali, M. High Selectivity of N-Doped ZnO Nano-Ribbons in Detecting H2, O2 and CO2 Molecules: Effect of Negative-Differential Resistance on Gas-Sensing. Sens. Actuators B Chem. 2018, 270, 167–178. [Google Scholar] [CrossRef]
  57. Karthik, T.V.K.; Hernández, A.G.; Kudriavtsev, Y.; Gómez-Pozos, H.; Ramírez-Cruz, M.G.; Martínez-Ayala, L.; Escobosa-Echvarria, A. Sprayed ZnO Thin Films for Gas Sensing: Effect of Substrate Temperature, Molarity and Precursor Solution. J. Mater. Sci. Mater. Electron. 2020, 31, 7470–7480. [Google Scholar] [CrossRef]
  58. Petrov, V.V.; Sysoev, V.V.; Starnikova, A.P.; Volkova, M.G.; Kalazhokov, Z.K.; Storozhenko, V.Y.; Khubezhov, S.A.; Bayan, E.M. Synthesis, Characterization and Gas Sensing Study of ZnO-SnO2 Nanocomposite Thin Films. Chemosensors 2021, 9, 124. [Google Scholar] [CrossRef]
  59. Çolak, H.; Karaköse, E. Synthesis and Characterization of Different Dopant (Ge, Nd, W)-Doped ZnO Nanorods and Their CO2 Gas Sensing Applications. Sens. Actuators B Chem. 2019, 296, 126629. [Google Scholar] [CrossRef]
  60. Shirage, P.M.; Rana, A.K.; Kumar, Y.; Sen, S.; Leonardi, S.G.; Neri, G. Sr- and Ni-Doping in ZnO Nanorods Synthesized by a Simple Wet Chemical Method as Excellent Materials for CO and CO2 Gas Sensing. RSC Adv. 2016, 6, 82733–82742. [Google Scholar] [CrossRef]
Figure 1. CO2 sensor fabrication process steps: (1) cleaning of Si wafer with piranha solution, (2) thermal oxidation, (3) negative lithography, (4) deposition of Ti/Au thin film, (5) lift-off process, (6) ZnO thin film deposition, and (7) final annealing (explanation of colors: light grey: Si wafer; dark grey: oxidized Si wafer; orange: Au, yellow: Ti/Au thin film, blue: ZnO thin film).
Figure 1. CO2 sensor fabrication process steps: (1) cleaning of Si wafer with piranha solution, (2) thermal oxidation, (3) negative lithography, (4) deposition of Ti/Au thin film, (5) lift-off process, (6) ZnO thin film deposition, and (7) final annealing (explanation of colors: light grey: Si wafer; dark grey: oxidized Si wafer; orange: Au, yellow: Ti/Au thin film, blue: ZnO thin film).
Sensors 25 02705 g001
Figure 2. XRD patterns of the SiO2 substrate and the pure and Na-doped ZnO thin films.
Figure 2. XRD patterns of the SiO2 substrate and the pure and Na-doped ZnO thin films.
Sensors 25 02705 g002
Figure 3. (a) Response, (b) response times, and (c) recovery times of pure ZnO, 2.5% Na-doped, 5% Na-doped, and 7.5% Na-doped sensing films.
Figure 3. (a) Response, (b) response times, and (c) recovery times of pure ZnO, 2.5% Na-doped, 5% Na-doped, and 7.5% Na-doped sensing films.
Sensors 25 02705 g003
Table 1. ZnO thin films manufactured for CO2 sensing: fabrication process, dopant concentrations, and sensing properties: optimum response and response/recovery time at CO2 concentration and temperature (T).
Table 1. ZnO thin films manufactured for CO2 sensing: fabrication process, dopant concentrations, and sensing properties: optimum response and response/recovery time at CO2 concentration and temperature (T).
Fabrication ProcessDopingCO2 Concentration Topt (°C)Sensing ResponseResponse/Recovery Time Ref.
direct current reactive magnetron sputtering-8.5 mbar1002.175 s/10 min[34]
successive chemical solution deposition 1% Al1000 ppm25~22.520 s/150 s[35]
chemical bath deposition 2600 ppm300~17%-[28]
direct current reactive magnetron sputtering -1000 ppm300 1.13%20 s/20 s[36]
sol-gel 5% In100 ppm200 16%-[31]
sol-gel 5% Ca2500 ppm45011310 s/10 s[37]
spray pyrolysis-1000 ppm200~6501.[38]
radio-frequency magnetron sputtering-15%300 9.97857 s/107 s[39]
aerosol spray pyrolysis0.1 at. % Mn100 ppm25 (under UV)66%1.[32]
sol-gel 2.5 at. % Na50 sccm2581.9%282.73 s/472.3 s[23]
sol-gel 3% Ca/1% Al5% CO2300 52.[11]
sol-gel 5% Ca2500 ppm450 ~703.[40]
chemical spray pyrolysis-400 ppm350 65%75 s/108 s[33]
chemical bath deposition-15%300 9~75 s/~150 s[41]
spray pyrolysis-100 Torr150~37-[42]
wet chemical synthesis5% Ca50,000 ppm350 53%-[43]
wet chemical synthesis5% Ca25,000 ppm350 ~32%-[21]
sol-gel 5% Al/1% Mg2000 ppm2001.9 -[30]
chemical bath deposition3% Cd100 ppm125 88.24%11 s/10 s[12]
wet chemical solution depositionZnO/CuO2500 ppm375 47% [44]
in-situ annealing2.4 at. % N500 ppm-5-[45]
sol-gel spin coating4.0 at. % La200 SCCM25122.71%24.4 s/44 s[19]
sol-gel spin coating5% Al100 ppm227100%90 s/160 s[20]
sol-gel spin coating-200 ppm450~15020 s/40 s[46]
Table 2. Morphology, film thickness, particle size distribution, mean particle size and mean film thickness of: (1) pure ZnO, (2) 2.5% Na-doped, (3) 5% Na-doped and (4) 7.5% Na-doped thin films. FE-SEM top view and cross section photos are presented in 2.5 kV voltage, x100,000 magnification, 3 mm working distance (WD), and 100 nm scale bar.
Table 2. Morphology, film thickness, particle size distribution, mean particle size and mean film thickness of: (1) pure ZnO, (2) 2.5% Na-doped, (3) 5% Na-doped and (4) 7.5% Na-doped thin films. FE-SEM top view and cross section photos are presented in 2.5 kV voltage, x100,000 magnification, 3 mm working distance (WD), and 100 nm scale bar.
Top ViewCross SectionSize DistributionMean Particle Size (nm)Mean Film Thickness (nm)
1Sensors 25 02705 i001Sensors 25 02705 i002Sensors 25 02705 i00356 a ± 16217 a ± 8
2Sensors 25 02705 i004Sensors 25 02705 i005Sensors 25 02705 i00657 a ± 19247 b ± 10
3Sensors 25 02705 i007Sensors 25 02705 i008Sensors 25 02705 i00959 a ± 16268 c ± 10
4Sensors 25 02705 i010Sensors 25 02705 i011Sensors 25 02705 i01260 a ± 20270 c ± 13
+/− shows the standard deviation between the replicates. Values not sharing the same superscript are significantly different (p < 0.05).
Table 3. AFM 2D and 3D topographies and RMS roughness of: (1) pure ZnO, (2) 2.5% Na-doped, (3) 5% Na-doped, and (4) 7.5% Na-doped thin films.
Table 3. AFM 2D and 3D topographies and RMS roughness of: (1) pure ZnO, (2) 2.5% Na-doped, (3) 5% Na-doped, and (4) 7.5% Na-doped thin films.
2D Topography3D TopographyRMS Roughness sq (nm)
1Sensors 25 02705 i013Sensors 25 02705 i01412.3
2Sensors 25 02705 i015Sensors 25 02705 i0167.7
3Sensors 25 02705 i017Sensors 25 02705 i0189.0
4Sensors 25 02705 i019Sensors 25 02705 i0209.2
Table 4. Sensing response represented as Net Signal current (A) over time and at varying CO2 concentrations with (1) pure ZnO, (2) 2.5% Na-doped, (3) 5% Na-doped, and (4) 7.5% Na-doped thin films.
Table 4. Sensing response represented as Net Signal current (A) over time and at varying CO2 concentrations with (1) pure ZnO, (2) 2.5% Na-doped, (3) 5% Na-doped, and (4) 7.5% Na-doped thin films.
Sensor Response (A)—Time (min)Sensor Response (A)—CO2 Concentration (ppm)
1Sensors 25 02705 i021Sensors 25 02705 i022
2Sensors 25 02705 i023Sensors 25 02705 i024
3Sensors 25 02705 i025Sensors 25 02705 i026
4Sensors 25 02705 i027Sensors 25 02705 i028
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Stramarkou, M.; Bardakas, A.; Krokida, M.; Tsamis, C. Fabrication of ZnO Thin Films Doped with Na at Different Percentages for Sensing CO2 in Small Quantities at Room Temperature. Sensors 2025, 25, 2705. https://doi.org/10.3390/s25092705

AMA Style

Stramarkou M, Bardakas A, Krokida M, Tsamis C. Fabrication of ZnO Thin Films Doped with Na at Different Percentages for Sensing CO2 in Small Quantities at Room Temperature. Sensors. 2025; 25(9):2705. https://doi.org/10.3390/s25092705

Chicago/Turabian Style

Stramarkou, Marina, Achilleas Bardakas, Magdalini Krokida, and Christos Tsamis. 2025. "Fabrication of ZnO Thin Films Doped with Na at Different Percentages for Sensing CO2 in Small Quantities at Room Temperature" Sensors 25, no. 9: 2705. https://doi.org/10.3390/s25092705

APA Style

Stramarkou, M., Bardakas, A., Krokida, M., & Tsamis, C. (2025). Fabrication of ZnO Thin Films Doped with Na at Different Percentages for Sensing CO2 in Small Quantities at Room Temperature. Sensors, 25(9), 2705. https://doi.org/10.3390/s25092705

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

Article Metrics

Back to TopTop