Next Article in Journal
Improved Hardness and Thermal Stability of Nanocrystalline Nickel Electrodeposited with the Addition of Cysteine
Previous Article in Journal
Nano-Bio Interaction between Blood Plasma Proteins and Water-Soluble Silicon Quantum Dots with Enabled Cellular Uptake and Minimal Cytotoxicity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanomaterial-Based CO2 Sensors

1
Department of Petroleum Engineering, Louisiana State University, Baton Rouge, LA 70803, USA
2
Department of Mechanical and Industrial Engineering, Louisiana State University, Baton Rouge, LA 70803, USA
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(11), 2251; https://doi.org/10.3390/nano10112251
Submission received: 16 October 2020 / Revised: 6 November 2020 / Accepted: 9 November 2020 / Published: 13 November 2020
(This article belongs to the Section Nanoelectronics, Nanosensors and Devices)

Abstract

:
The detection of carbon dioxide (CO2) is critical for environmental monitoring, chemical safety control, and many industrial applications. The manifold application fields as well as the huge range of CO2 concentration to be measured make CO2 sensing a challenging task. Thus, the ability to reliably and quantitatively detect carbon dioxide requires vastly improved materials and approaches that can work under different environmental conditions. Due to their unique favorable chemical, optical, physical, and electrical properties, nanomaterials are considered state-of-the-art sensing materials. This mini-review documents the advancement of nanomaterial-based CO2 sensors in the last two decades and discusses their strengths, weaknesses, and major applications. The use of nanomaterials for CO2 sensing offers several improvements in terms of selectivity, sensitivity, response time, and detection, demonstrating the advantage of using nanomaterials for developing high-performance CO2 sensors. Anticipated future trends in the area of nanomaterial-based CO2 sensors are also discussed in light of the existing limitations.

1. Introduction

Carbon dioxide (CO2) is vital to life on Earth. It is also pivotal for many biological and industrial processes. The concentration of CO2 in Earth’s atmosphere is currently close to 412 parts per million (ppm) which represents a 47% increase since the beginning of the Industrial Age, when the concentration was near 280 ppm, and an 11% increase since 2000, when it was near 370 ppm [1]. The rising concentration of CO2 in the atmosphere is driving up surface temperatures and causing ocean acidification, making it one of the primary climate change contributors [2]. Moreover, extended exposure to high CO2 concentration can also be lethal for human beings. Thus, sensing and monitoring of CO2 is fundamental to gaining knowledge about CO2-affected mechanisms and controlling them.
The conditions in which CO2 is monitored dictate the method and the materials that can be used in the sensor [3]. A variety of sensors have been developed based on different sensing principles including optical absorption [4,5,6], electrical resistance [7,8,9,10], field-effect transistors [11,12,13], and amperometry [14]. However, conventional CO2 sensors have several limitations such as higher cost, heavier weight, bigger size, and less durability [15]. To enable the widespread adoption of CO2 sensors in many aspects of modern society, inexpensive mass-producible devices are needed that offer simplicity, robustness, and ultralow power demand. Furthermore, there is a need for an accurate and reliable sensor that operates at harsh conditions of pressure and temperature such as in deep oil wells or nuclear reactors. To meet these needs, there has been a growing interest in recent years in nanomaterial-based CO2 sensors due to their cost efficiency, durability in harsh conditions, and stability [16,17]. The use of nanomaterials has been shown to improve traditional sensing techniques in terms of sensitivity, stability, response time, and selectivity [18]. The aforementioned benefits of nanomaterials are often ascribed to their high surface to volume ratio, high surface energy, quantum size effect, and high electron conductivity [19].
Nanomaterials used for sensing may consist of organic, inorganic or hybrid components. In CO2 sensing, it is desirable to use materials with functional groups that display application interaction. Organic materials exhibit many desirable properties such as mass transport, chemical reactivity, and gas diffusivity in the context of CO2 sensing. On the other hand, inorganic materials show mechanical stability, conductivity, and optical properties [20]. Hybrid materials have also been explored for CO2 sensing in which unique chemical conjugates of organic and inorganic components are brought together by specific interactions for synergistic improvement of their functional properties [21].
The main thrust of this mini-review is to document the most significant work done in the area of nanomaterials-based sensing of CO2, primarily in the free gas state, in the last two decades. On the basis of the key sensing mechanism, CO2 nanosensors can be categorized based on electrochemical principle (which includes chemiresistive, capacitive, and inductive sensors), and based on optical sensing (which includes surface plasmonic resonance, colorimetric, refractometric, and non-dispersive infrared sensors). These broad categories of sensors are discussed in the following sections in terms of their sensing mechanisms and the most significant nanomaterials used. Their common applications, inherent drawbacks, and key performance metrics are compared to serve as a guide for selecting the CO2 sensor appropriate for meeting the application-specific demands and requirements.
The manifold application fields as well as the huge range of CO2 concentrations to be measured make CO2 sensing a challenging task. In view of the great diversity of CO2 measurement tasks, this mini-review article is not intended to provide solutions for all possible measurement problems. Rather, its scope is restricted to surveying the current developments in the field of nanomaterial-based CO2 sensing, highlighting their strengths, weaknesses and the most representative application space. Anticipated future trends in the area of CO2 sensing are also discussed in light of the aforementioned comparison.

2. Electrochemical CO2 Sensors

The main working principle of these types of sensor is the variation in electrical properties upon chemical interaction with CO2. The most commonly used gas sensors in this category are chemiresistive which measure the change in resistance, and capacitive, which measure the change in capacitance corresponding to the change in the CO2 concentration [22]. The simplicity of this sensing technique as well as its cost-effectiveness were the main incentives in its wide development. The most widely used type of nanomaterials in this area are inorganic metal oxides (such as ZnO, SnO2, CuO, and CdO) where p-n-type can act as the base material for the sensing layer. When CO2 comes in contact with the semiconductor nanomaterial layer a surface interaction may occur through oxidation/reduction, electron charge transfer, adsorption, or chemical reaction [23]. The chemical interaction of the adsorbate (CO2) and adsorbent semiconducting nanomaterial causes a charge depletion layer with upward bending energy bands that lead to change in electrical properties [24,25]. This is represented in the schematic in Figure 1.
It has been reported that the sensitivity of such sensors mostly depends upon the nanostructure’s porosity [26,27,28]. At dry air conditions, the oxygen species chemisorbed on the surface of the sensing nanomaterial causes an electron depletion layer which creates a Schottky barrier [29,30,31]. The interaction between CO2 and chemisorbed oxygen species gas off carbon monoxide and electrons. Subsequently, this reaction decreases the Schottky surface barrier which increases conductivity. CO2, which is normally regarded as reducing gas except in some work it has been reported as oxidizing gas [32,33,34], increases the resistivity upon interaction with oxygen species [35]. Furthermore, it has been observed that high surface area and increased uniform pore size distribution make a great contribution to sensitivity as well as response time [36]. The sensor response (S.R) is measured as a function of change in resistance as shown by the equations below [37,38]:
Reducing   gases   S . R   =   R a   R g R g Oxidizing   gases   S . R   =   R g   R a R a
where Rg is the resistance of the sensitive nanomaterial layer in the presence of CO2 while Ra is the resistance of the sensing layer in air. Studies have shown that the high surface to volume ratio of nanomaterials yields a quicker response time [39,40]. Similarly, the presence of an oxygen-rich background while sensing CO2 concentration increases the sensor’s response while the presence of amine functional groups can hasten the recovery time at varying temperatures [36,41,42]. CO2 as a strong Lewis acid strongly tends to accept additional pairs of electrons from primary and secondary amines that are strong Lewis bases. The aforementioned reaction is a covalent bond that occurs due to chemisorption of CO2 and the sensing nanomaterial. In assessing cross-sensitivity to identify the sensors’ specificity, some works have demonstrated the dependence of CO2 sensing on ambient humidity [43,44]. CO2 is known to hinder proton diffusion by blocking Grotthus pathways in the sensing nanomaterial layer [45,46]. These aforementioned phenomena cause less proton diffusion, and increased charge transfer resistance contributes to the overall sensitivity.
An additional parameter that has enhanced CO2 sensitivity is the metal oxide doping. Based on the choice of material, if the sensing nanomaterial and the dopant can create a p–n junction, this will generally increase the resistance baseline. Furthermore, increasing doping from 0% to a specific threshold can diminish both response and recovery times. However, surpassing this threshold can have counter-effects in terms of sensitivity, response, and recovery times [7]. Some of the studies conducted CO2 sensing experiments in the presence of oxygen concentrations in the atmospheric background. However, in other published works [42,47] the sensitivity of nanomaterials to CO2 was not affected in the absence of oxygen. Similarly, the effect of humidity on the sensing nanomaterial was identified as debatable based on the nanomaterials’ selection and other factors. In some cases, humidity acted as a sensor response enhancer due to favorable CO2 adsorption [48]. In other cases, the inhibited sensor performance was attributed to the consumption of huge number of oxygen ions. This debatable humidity effect is illustrated in Figure 2 based on published studies.
Despite the varying temperatures at which CO2 sensing took place, no chemiresistive sensor was investigated beyond 450 °C in dry air [50]. The application of this type of sensor has been extended in different fields including food processing and agriculture industry [51], medical diagnosis [52], and environmental monitoring [53,54]. Table 1 summarizes the most recent and significant work undertaken in nanomaterial-based chemiresistive and capacitive CO2 sensors.

3. Optical Sensors

3.1. Surface Plasmon Resonance (SPR) Sensors

The surface plasmon resonance (SPR) effect takes place when incident light, at a particular angle of incidence, causes resonant oscillation of electrons at the interface of two media [60]. The high sensitivity of the angle of incidence to the refractive index alterations is the basis of CO2 sensing in terms of adsorption and desorption. The changes in the effective refractive index of the sensing nanomaterial indicate a change in the concertation of CO2. In other words, these phenomena could be explained by what is referred to as the Kretschmann configuration. A few materials that exhibit negative real and small imaginary dielectric constant (Cu, Al, Au, and Ag) are the only materials showing surface plasmon capability [61]. The basis of this technique is the interaction of nanomaterial with the incident light. Owing to their superior optical properties, the absorption of incident light in nanomaterials can result in enhanced localized electric field at the localized surface plasmon resonance (LSPR) frequency. On metallic nanoparticles at the quantum size the conduction band is discretized which further enhances charge transfer [62]. In this detection scheme, a light source is used to illuminate a plasmonic nanomaterial. A portion of the light is reflected off the surface of the plasmonic nanomaterial [63,64] while a portion of light is absorbed by the surface electrons at a unique angle called the angle of resonance. This results in electronic resonance where they are called surface plasmons [65]. Surface plasmon resonance is a condition that alters the dielectric constant adjacent to the nanomaterial’s layer [66,67,68]. The sensitivity of surface plasmons results in a loss in the intensity of the reflected beam. The location and the shape of the observed dip in the SPR reflection intensity curve can be used to conclude information about the sensor’s surface. The kinetics of CO2 binding with the sensing material can also be studied through a time-resolved SPR response curve. In this curve, as CO2 interacts with the sensing material, the response increases as the binding of CO2 with the sensing material increases and once the system reaches equilibrium, CO2 starts unbinding or dissociating. The overall sensing mechanism is illustrated in the schematic in Figure 3.
The nanomaterials most used for plasmonic sensing of CO2 are carbon nanotubes (CNTs). The main reason for this is that the high selectivity of CNTs to CO2 at room temperature among other gases was ascribed to the high affinity of CNTs to CO2 causing high electron density and hole depletion [70,71,72]. However, the use of CNTs as plasmonic or optical sensors is still limited because they have an excitation regime in the infrared (IR) and ultraviolet (UV) regions and cross-sensitivity. In addition, some works have indicated that a need to improve reversibility and selectivity when sensing CO2 at room temperature [73]. Since this type of sensor usually shows non-toxicity with high sensitivity at low cost, it is widely used in biosensing application and pharmaceutical analysis [74,75].

3.2. Colorimetric Sensing

Colorimetric CO2 sensing is a technique in which the sensing material (such as pH sensitive dyes or quantum dots) exhibits a color change upon chemical reaction or adsorption of CO2. Colorimetric sensors have been commonly used as semi-quantitative or qualitative non-invasive sensing technique in biological applications [76]. Their application is also reported for soil evaluation in carbon capture and storage sites [77]. The sensing technique is based on visual detection of the change in color due to a change of intensity at a specific wavelength on a dye. Depending on the material used, this sensing technique can be fully reversible or non-reversible [76,78]. In some cases, the Beer–Lambert law has been used to report differences in gas concentrations [79]. The aforementioned sensing technique is represented in Figure 4, where the pH change is depicted in response to CO2 with specific intensity.
The colorimetric nanomaterials sensing used to detect CO2 so far have been either complementary with fiber optic sensing or with dyes. Chu et al. [80] investigated the use of coarse silica nanoparticles (200 nm) as a phase transfer agent in fiber optic sensor. It was concluded that their proposed sensor had higher sensitivity and better linearity with slower response time compared to similar sensors. The enhanced sensitivity was ascribed to the high surface-to-volume ratio of silica nanoparticles.
Quantum dots (QD) have been receiving much attention in gas sensing during the last decade. Specifically, in the area of colorimetric sensing, QDs are favorable in comparison to dye molecules owing to their quantum confinement as well as immunity to electromagnetic interference, narrow emission, wavelength size dependency, and superior photostability [81]. QDs are semiconductor nanocrystals (few nanometers in size) in which quantum confinement effects are proved. In other words, they are a group of nanocrystals in which electrons and electron-hole pairs are tightly confined. Quantum dots are either synthesized from semi-conductor compounds such as PbSe, CdSe, or PbS or they can be made simply of a single element such as germanium or silicon [82,83,84]. QDs are known for their unique optical and plasmonic properties due to their increased energy band gap as size decreases. Due to the energy quantization effects, quantum dots are capable of emitting colors more accurately. Upon reaction with CO2 and pH change in the presence of light, a photon hits an electron at the valence band. The excited electron moves from the valence band to the conduction band then relaxes again to a lower energy band by releasing a photon that matches the energy difference. This difference in energy corresponds to the color shift. The color shift would be affected by both the chemical reaction as well as the size of the quantum dot. Larger dots have smaller energy between bands with longer wavelengths and vice versa, as illustrated in Figure 5. The CO2 layer adsorbed on the surface of the QD will have an impact on the bandgap. However, the nature of this effect depends on the material of the QD and the type of interaction occurring with CO2. Since QDs have not been widely used in CO2 sensing it is hard to claim whether it would enhance the band gap or not. Many things could happen upon the adsorption of CO2 on the QDs’ surface. Firstly, the adsorption can cause a change in the intensity of the resonance. Secondly, Jin et al., 2006 [85] showed that the presence of a new layer of a different material on the surface of QD can cause a red shift due to a decrease in the bandgap. On the other hand, other researchers such as Joo et al., 2018 [86] proved that the adsorption of other chemical groups can cause surface defect passivation. This can result in a blue shift due to increasing bandgap. Furthermore, other researchers including Saha and Sarkar, 2014 [87] observed a bowing effect while increasing the concentration of the sorbed layer on the surface of QD. The effect of organic as well as inorganic passivation of QDs has been reviewed [88] with no sufficient referencing to the type of interaction of CO2 with QD.

3.3. Refractometric Fiber-Optic Sensors

The use of fiber-optic CO2 sensing is quite prominent in environmental monitoring as well as aquaculture industry applications [89]. Fiber-optic sensing can be integrated with a variety of sensing platforms such as SPR sensing [90,91]. In SPR fiber optics, the core acts as the prism in the Kretschmann configuration. The same plasmon excitation mechanism is applied as in the prism-based setting. At a specific interval, the cladding is etched and the core is coated with the plasmonic nanomaterial [67]. A plethora of work has been done in CO2 interaction of nanomaterial coated over an unclad or clad etched fiber core causing a variation in the refractive index of the sensing material [92]. However, there are some drawbacks to SPR fiber optics including the integration of SPR into a multiplexed platform and challenges related to the limit of detection [93]. Shivananju et al. [94] developed a clad-etched fiber Bragg grating (FBG) with polyallylamine-amino-carbon nanotube coated on the surface of the core for detecting the concentrations of CO2 gas at room temperature over a wide range of concentrations (1000–4000 ppm). A reversible and reproducible linear response of Bragg wavelength shift was observed at a limit of detection (LOD) of 75 ppm. On the other hand, it was possible to make use of a metal-organic framework (MOFs) nano-porous structure in the CO2 optical-fiber sensor [92]. The main sensing material used was Cu-BTC (Cu-benzene-1,3,5-tricarboxylate) was coated over unclad core single-mode optical fiber. At the tested sensing region (5 cm long), this material recorded LOD of 20 ppm and a figure of merit (FOM) of 100 ppm. Reversibility investigation for the aforementioned material was not investigated [95]. In addition, other materials such as CNTs were coated on the surface of optical fibers and results were compared with an ordinary CNT uncoated optic fiber. The results proved enhanced selectivity and sensitivity due to the increased surface to volume ratio and CO2 reduced activation energy when interacting with CNTs [96]. A generalized simple schematic of optical fiber-based CO2 sensing using nanomaterials is represented in Figure 6.
Fiber optic-based gas sensors evaluates the concentration of gas as a function of the change in the refractive index (RI) [97]. Upon detection of CO2 concentration and changes in RI, frequency shifts of optical resonance are observed. Some of the drawbacks noted for these types of sensor are non-reliability, poor selectivity, and contamination issues. In recent work in this field, the aforementioned technique was merged with nanomaterials and used under the same umbrella. This work was motivated by the need for remote sensing, lessened noise interference, and a good candidate for CO2 sensing at harsh conditions [98]. Generally, the cladding of the fiber is etched using HF (hydrofluoric acid) to expose the core of the optical fiber. Then a sol or a thin film is coated in the etched place preferentially with porous nanostructures film. The increased surface to volume ratio as well as high porosity of nanomaterials help to increase the sensitivity.

3.4. Non-Dispersive Infrared Sensors

Optical detection based on the non-dispersive infrared (NDIR) method is a well-known and established concept for detecting gases. It is very common to use NDIR in monitoring the indoor air quality as well as automotive applications [99,100]. This sensing technique makes use of shining an infrared (IR) source on a sample in a chamber with an optical filter and an infrared detector, as shown in the schematic in Figure 7. Carbon dioxide can be easily identified by infrared spectroscopy because it is an infrared active molecule, which absorbs a 4.24 µm wavelength [15]. The light passes through the optical filter to obtain shorter wavelengths that are attenuated and longer wavelengths that are transmitted. If the CO2 is below the detection limit or absent, the detector will match the intensity to the reference level [100,101]. The IR intensity at the detector will reduce in accordance with an exponential relationship known as the Beer–Lambert Law: I = I 0 e k P , where I is the intensity at the detector subsequent to optical filtering, I0 is the initial intensity prior to interacting with gas, k and P are the absorption coefficient and the gas concentration, respectively [102,103].
A few studies have reported the use of nanomaterials to improve either the IR emitting source [104] or the photodetectors [105]. Muller et al. [104] showed that IR emitter integrated with Pt-on-Si-needles demonstrated a 2.6 times higher IR emission without wavelength-dependent interference patterns as compared to an uncoated Si-based emitter at the same membrane temperature. Koppens et al. [105] evaluated state-of-the-art photodetectors based on graphene and other nanomaterials such as plasmonic nanoparticles. Similarly, Pusch et al. [106] investigated replacing the ordinary standard thermal emitter with CMOS nanoplasmonic tungsten crystal. They discovered that the sensitivity and signal to noise ratio to CO2 increased by 400%. This enhancement was attributed to slow-wave plasmonic lattice resonance in addition to elevated plasmon to light coupling.
Although this type of sensor has high accuracy, allows fast measurements and has a good long-term stability, even after decades of optimization, some inherent drawbacks of its optoelectronic mode of operation remain unresolved such as device complexity, power consumption, scalability, and cost [107,108]. Furthermore, the detection limit and sensitivity are highly dependent on the power-related parameters such as light intensity. A key issue that has not been well addressed in the literature is the efficacy of CO2 selectivity among the interference of adsorption bands [109]. Table 2 summarizes some of the recent significant work done in optical CO2 sensors using nanomaterials.

4. Challenges and Anticipated Future Trends

Owing to their unique properties, nanomaterials are emerging as key players in improving CO2 sensing technology. However, many challenges remain in catering to the diverse sensing demands. For example, CO2 selectivity remains problematic in the different types of CO2 sensor discussed here. The main concept through which CO2 selectivity is obtained nowadays is through the use of different gases’ working temperatures. In other words, at a specific range of temperature, one of the gases gives a higher response as compared to other gases due to chemical or physical interactions on the surface of the nanomaterial. This “lock and key” model is a widely used technique in biological fields where a sensing material has high specificity to lock on to the intended measurand. There is also a need for durable, low-maintenance, real-time sensors particularly in harsh environments, such as in a borehole and subsea. Thus, future work in nanomaterials should stack up that need. Multi-purpose sensing is also a fast-developing area, where the same sensing element is used for measuring multiple parameters. Some applications with high temperature requirements, such as in nuclear reactors would require sensors that can operate in range of 600–700 °C. More work in the future is required to enhance the temperature tolerance of these nanosensors. However, the selection of the elements in these composites seems non-systematic that is based on haphazard selection and trial and error. Therefore, there is a need for more basic studies and modelling to conceptualize the interaction of CO2 with different materials’ families.
Nanomaterial CO2 sensing is an interdisciplinary field that necessitates the identification of several parameters before deciding which nanosensor type to use. These parameters include but are not limited to dynamic sensing range, temperature, humidity, response, and reversibility. In addition, some other considerations such as size and cost should also be considered for successful CO2 sensing. A marketplace comparison of the commercially available CO2 sensors reveals useful insights [118]. Firstly, electrochemical sensors are typically less durable than optical sensors. In addition, more work should be done in this field when it comes to cross-sensitivity. While, chemiresistive CO2 sensors continue to be popular for a variety of applications due to their easy handling and greater tolerance to humidity and temperature changes. Secondly, the use of colorimetric and SPR is not as common as electrochemical ones due to the reduced accuracy in quantitative measurements. Thirdly, the use of refractometric CO2 sensing appears to be an emerging field with great potential since it can reach inaccessible areas and still measure with high accuracy. Multiplexing remains a topic that needs to be addressed in this field to make it more useful and achieve maximum performance for its cost. Finally, NDIR is a long-lasting type of sensor and it can measure CO2 at high concentrations with good selectivity. However, it can be affected by temperature and humidity. The authors think that nanomaterials have demonstrated tremendous potential in improving CO2 sensing towards higher efficiency and accuracy and although opportunities of improvement remain, nanomaterial-based sensing will continue to be a growing research priority.

5. Conclusions

This review summarizes the most significant work done in the area of nanomaterial-based CO2 sensing in the last two decades. The main sensing techniques were categorized by their sensing mechanism under two main categories and the role of nanomaterials in each sensing technique was highlighted. Their common applications, inherent drawbacks, and key performance metrics are compared to serve as a guide for selecting the CO2 sensor appropriate for meeting the application-specific demands and requirements.

Author Contributions

Conceptualization, J.S.; visualization, M.Y.R.; writing—Original draft preparation, M.Y.R. and J.S.; writing— Review and editing, M.R.G.; administration and supervision, J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Möller, D.; Decker, M.; Zosel, J.; Oelßner, W. Carbon dioxide in general. In Carbon Dioxide Sensing Fundamentals, Principles, and Applications; Wiley-VCH: Weinheim, Germany, 2019; pp. 39–40. [Google Scholar]
  2. CO2 and Ocean Acidification: Causes, Impacts, Solutions. Available online: https://www.ucsusa.org/resources/co2-and-ocean-acidification (accessed on 9 July 2020).
  3. Bezzon, V.D.N.; Montanheiro, T.L.A.; Menezes, B.R.C.D.; Ribas, R.G.; Righetti, V.A.N.; Rodrigues, K.F.; Thim, G.P. Carbon Nanostructure-based Sensors: A Brief Review on Recent Advances. Adv. Mater. Sci. Eng. 2019, 2019, 1–21. [Google Scholar] [CrossRef] [Green Version]
  4. Starecki, F.; Charpentier, F.; Doualan, J.-L.; Quetel, L.; Michel, K.; Chahal, R.; Troles, J.; Bureau, B.; Braud, A.; Camy, P.; et al. Mid-IR optical sensor for CO2 detection based on fluorescence absorbance of Dy3+: Ga5Ge20Sb10S65 fibers. Sens. Actuators B Chem. 2015, 207, 518–525. [Google Scholar] [CrossRef]
  5. Liu, L.; Morgan, S.P.; Correia, R.; Lee, S.-W.; Korposh, S. Multi-Parameter Optical Fiber Sensing of Gaseous Ammonia and Carbon Dioxide. J. Light. Technol. 2020, 38, 2037–2045. [Google Scholar] [CrossRef]
  6. Starecki, F.; Braud, A.; Doualan, J.-L.; Ari, J.; Boussard-Plédel, C.; Michel, K.; Nazabal, V.; Camy, P. All-optical carbon dioxide remote sensing using rare earth doped chalcogenide fibers. Opt. Lasers Eng. 2019, 122, 328–334. [Google Scholar] [CrossRef]
  7. Xiong, Y.; Xue, Q.; Ling, C.; Lu, W.; Ding, D.; Zhu, L.; Li, X. Effective CO2 detection based on LaOCl-doped SnO2 nanofibers: Insight into the role of oxygen in carrier gas. Sens. Actuators B Chem. 2017, 241, 725–734. [Google Scholar] [CrossRef]
  8. George, R.; Kumar, L.A.; Alagappan, M. Synthesis of nanotubular NiO-CNT composite and its application in temperature independent CO2 gas sensors fabricated using interdigitated silver electrode. Dig. J. Nanomater. Biostruct. 2019, 14, 213–224. [Google Scholar]
  9. Hammad, A.B.A.; Elnahrawy, A.M.; Youssef, A.M.; Youssef, A.M. Sol gel synthesis of hybrid chitosan/calcium aluminosilicate nanocomposite membranes and its application as support for CO2 sensor. Int. J. Biol. Macromol. 2019, 125, 503–509. [Google Scholar] [CrossRef]
  10. Stassen, I.; Dou, J.-H.; Hendon, C.; Dincă, M. Chemiresistive Sensing of Ambient CO2 by an Autogenously Hydrated Cu3(hexaiminobenzene)2 Framework. ACS Cent. Sci. 2019, 5, 1425–1431. [Google Scholar] [CrossRef] [Green Version]
  11. Ersöz, B.; Schmitt, K.; Wöllenstein, J. Electrolyte-gated transistor for CO2 gas detection at room temperature. Sens. Actuators B Chem. 2020, 317, 128201. [Google Scholar] [CrossRef]
  12. Santonico, M.; Zompanti, A.; Sabatini, A.; Vollero, L.; Grasso, S.; Mezza, C.D.; Pennazza, G. CO2 and O2 Detection by Electric Field Sensors. Sensors 2020, 20, 668. [Google Scholar] [CrossRef] [Green Version]
  13. Lv, A.; Pan, Y.; Chi, L. Gas Sensors Based on Polymer Field-Effect Transistors. Sensors 2017, 17, 213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Huang, X.-J.; Aldous, L.; O’Mahony, A.M.; Campo, F.J.D.; Compton, R.G. Toward Membrane-Free Amperometric Gas Sensors: A Microelectrode Array Approach. Anal. Chem. 2010, 82, 5238–5245. [Google Scholar] [CrossRef] [PubMed]
  15. Boudaden, J.; Klumpp, A.; Hecker, C.; Eisele, I.; Joseph, Y. Functionalized nanoparticles for CO2 sensor. Funct. Nanostruct. Proc. 2018, 64–66. [Google Scholar]
  16. Deep, A.; Kumar, S.; Elsevier, S. Advances in Nanosensors for Biological and Environmental Analysis; Elsevier: St. Louis, MO, USA, 2019. [Google Scholar]
  17. Abdel-Karim, R.; Reda, Y.; Abdel-Fattah, A. Review—Nanostructured Materials-Based Nanosensors. J. Electrochem. Soc. 2020, 167, 037554. [Google Scholar] [CrossRef]
  18. Malik, R.; Tomer, V.K.; Mishra, Y.K.; Lin, L. Functional gas sensing nanomaterials: A panoramic view. Appl. Phys. Rev. 2020, 7, 021301. [Google Scholar] [CrossRef] [Green Version]
  19. Ge, L.; Mu, X.; Tian, G.; Huang, Q.; Ahmed, J.; Hu, Z. Current Applications of Gas Sensor Based on 2-D Nanomaterial: A Mini Review. Front. Chem. 2019, 7, 839. [Google Scholar] [CrossRef]
  20. Rebber, M.; Willa, C.; Koziej, D. Organic–inorganic hybrids for CO2 sensing, separation and conversion. Nanoscale Horiz. 2020, 5, 431–453. [Google Scholar] [CrossRef]
  21. Ananikov, V.P. Organic–Inorganic Hybrid Nanomaterials. Nanomaterials 2019, 9, 1197. [Google Scholar] [CrossRef] [Green Version]
  22. Fleischer, M. Advances in application potential of adsorptive-type solid state gas sensors: High-temperature semiconducting oxides and ambient temperature GasFET devices. Meas. Sci. Technol. 2008, 19, 042001. [Google Scholar] [CrossRef]
  23. Korotcenkov, G. Metal oxides for solid-state gas sensors: What determines our choice? Mater. Sci. Eng. B 2007, 139, 1–23. [Google Scholar] [CrossRef]
  24. Lin, T.; Lv, X.; Hu, Z.; Xu, A.; Feng, C. Semiconductor Metal Oxides as Chemoresistive Sensors for Detecting Volatile Organic Compounds. Sensors 2019, 19, 233. [Google Scholar] [CrossRef] [Green Version]
  25. Capone, S.; Forleo, A.; Francioso, L.; Rella, R.; Siciliano, P.; Spadavecchia, J.; Presicce, D.S.; Taurino, A.M. Solid State Gas Sensors: State of the Art and Future Activities. ChemInform 2004, 35. [Google Scholar] [CrossRef]
  26. Wang, L.; Lou, Z.; Fei, T.; Zhang, T. Enhanced acetone sensing performances of hierarchical hollow Au-loaded NiO hybrid structures. Sens. Actuators B Chem. 2012, 161, 178–183. [Google Scholar] [CrossRef]
  27. Zhang, H.; Zhu, Q.; Zhang, Y.; Wang, Y.; Zhao, L.; Yu, B. One-Pot Synthesis and Hierarchical Assembly of Hollow Cu2O Microspheres with Nanocrystals-Composed Porous Multishell and Their Gas-Sensing Properties. Adv. Funct. Mater. 2007, 17, 2766–2771. [Google Scholar] [CrossRef]
  28. Wang, L.; Zhou, T.; Zhang, R.; Lou, Z.; Deng, J.; Zhang, T. Comparison of toluene sensing performances of zinc stannate with different morphology-based gas sensors. Sens. Actuators B Chem. 2016, 227, 448–455. [Google Scholar] [CrossRef]
  29. Shinde, P.V.; Shinde, N.M.; Shaikh, S.F.; Lee, D.; Yun, J.M.; Woo, L.J.; Al-Enizi, A.M.; Mane, R.S.; Kim, K.H. Room-temperature synthesis and CO2-gas sensitivity of bismuth oxide nanosensors. RSC Adv. 2020, 10, 17217–17227. [Google Scholar] [CrossRef]
  30. Lin, T.; Lv, X.; Li, S.; Wang, Q. The Morphologies of the Semiconductor Oxides and Their Gas-Sensing Properties. Sensors 2017, 17, 2779. [Google Scholar] [CrossRef] [Green Version]
  31. Mirzaei, A.; Hashemi, B.; Janghorban, K. α-Fe2O3 based nanomaterials as gas sensors. J. Mater. Sci. Mater. Electron. 2015, 27, 3109–3144. [Google Scholar] [CrossRef]
  32. Son, M.; Pak, Y.; Chee, S.-S.; Auxilia, F.M.; Kim, K.; Lee, B.-K.; Lee, S.; Kang, S.K.; Lee, C.; Lee, J.S.; et al. Charge transfer in graphene/polymer interfaces for CO2 detection. Nano Res. 2018, 11, 3529–3536. [Google Scholar] [CrossRef]
  33. Abdelmounaïm, C.; Amara, Z.; Maha, A.; Mustapha, D. Effects of molarity on structural, optical, morphological and CO2 gas sensing properties of nanostructured copper oxide films deposited by spray pyrolysis. Mater. Sci. Semicond. Process. 2016, 43, 214–221. [Google Scholar] [CrossRef]
  34. Li, Y.; Li, G.; Wang, X.; Zhu, Z.; Ma, H.; Zhang, T.; Jin, J. Poly(ionic liquid)-wrapped single-walled carbon nanotubes for sub-ppb detection of CO2. Chem. Commun. 2012, 48, 8222. [Google Scholar] [CrossRef] [PubMed]
  35. Bhande, S.S.; Mane, R.S.; Ghule, A.V.; Han, S.-H. A bismuth oxide nanoplate-based carbon dioxide gas sensor. Scr. Mater. 2011, 65, 1081–1084. [Google Scholar] [CrossRef]
  36. Boudaden, J.; Klumpp, A.; Endres, H.-E.; Eisele, I. Towards Low Cost and Low Temperature Capacitive CO2 Sensors Based on Amine Functionalized Silica Nanoparticles. Nanomaterials 2019, 9, 1097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Viswanathan, P.; Patel, A.K.; Pawar, J.; Patwardhan, A.; Henry, R. Fabrication of Tin Oxide Nanoparticles for CO2 Gas Sensing Layer. IETE J. Res. 2018, 66, 460–465. [Google Scholar] [CrossRef]
  38. Kim, K.-M.; Jeong, H.-M.; Kim, H.-R.; Choi, K.-I.; Kim, H.-J.; Lee, J.-H. Selective Detection of NO2 Using Cr-Doped CuO Nanorods. Sensors 2012, 12, 8013–8025. [Google Scholar] [CrossRef] [PubMed]
  39. Wang, Y.; Wang, S.; Zhang, H.; Gao, X.; Yang, J.; Wang, L. Brookite TiO2 decorated α-Fe2O3 nanoheterostructures with rod morphologies for gas sensor application. J. Mater. Chem. A 2014, 2, 7935. [Google Scholar] [CrossRef]
  40. Lee, J.-H. Gas sensors using hierarchical and hollow oxide nanostructures: Overview. Sens. Actuators B Chem. 2009, 140, 319–336. [Google Scholar] [CrossRef]
  41. Boudaden, J.; Klumpp, A.; Endres, H.-E.; Eisele, I. Capacitive CO2 Sensor. Proceedings 2017, 1, 472. [Google Scholar] [CrossRef] [Green Version]
  42. Zhou, Y.; Jiang, Y.; Xie, T.; Tai, H.; Xie, G. A novel sensing mechanism for resistive gas sensors based on layered reduced graphene oxide thin films at room temperature. Sens. Actuators B Chem. 2014, 203, 135–142. [Google Scholar] [CrossRef]
  43. Lorek, A.; Majewski, J. Humidity Measurement in Carbon Dioxide with Capacitive Humidity Sensors at Low Temperature and Pressure. Sensors 2018, 18, 2615. [Google Scholar] [CrossRef] [Green Version]
  44. Wei, P.; Ning, Z.; Ye, S.; Sun, L.; Yang, F.; Wong, K.; Westerdahl, D.; Louie, P. Impact Analysis of Temperature and Humidity Conditions on Electrochemical Sensor Response in Ambient Air Quality Monitoring. Sensors 2018, 18, 59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Agmon, N. Mechanism of hydroxide mobility. Chem. Phys. Lett. 2000, 319, 247–252. [Google Scholar] [CrossRef]
  46. Kreuer, K.-D. Proton Conductivity: Materials and Applications. Chem. Mater. 1996, 8, 610–641. [Google Scholar] [CrossRef]
  47. Fan, K.; Qin, H.; Wang, L.; Ju, L.; Hu, J. CO2 gas sensors based on La1−xSrxFeO3 nanocrystalline powders. Sens. Actuators B Chem. 2013, 177, 265–269. [Google Scholar] [CrossRef]
  48. Marsal, A.; Dezanneau, G.; Cornet, A.; Morante, J. A new CO2 gas sensing material. Sens. Actuators B Chem. 2003, 95, 266–270. [Google Scholar] [CrossRef]
  49. Kim, D.Y.; Kang, H.; Choi, N.-J.; Park, K.H.; Lee, H.-K. A carbon dioxide gas sensor based on cobalt oxide containing barium carbonate. Sens. Actuators B Chem. 2017, 248, 987–992. [Google Scholar] [CrossRef]
  50. Dhahri, R.; Hjiri, M.; Mir, L.E.; Fazio, E.; Neri, F.; Barreca, F.; Donato, N.; Bonavita, A.; Leonardi, S.G.; Neri, G. ZnO:Ca nanopowders with enhanced CO2sensing properties. J. Phys. D Appl. Phys. 2015, 48, 255503. [Google Scholar] [CrossRef]
  51. Neethirajan, S.; Jayas, D.S.; Sadistap, S. Carbon Dioxide (CO2) Sensors for the Agri-food Industry—A Review. Food Bioprocess Technol. 2008, 2, 115–121. [Google Scholar] [CrossRef]
  52. Tricoli, A.; Nasiri, N.; De, S. Wearable and Miniaturized Sensor Technologies for Personalized and Preventive Medicine. Adv. Funct. Mater. 2017, 27, 1605271. [Google Scholar] [CrossRef]
  53. Fine, G.F.; Cavanagh, L.M.; Afonja, A.; Binions, R. Metal Oxide Semi-Conductor Gas Sensors in Environmental Monitoring. Sensors 2010, 10, 5469–5502. [Google Scholar] [CrossRef] [Green Version]
  54. Martinelli, G.; Carotta, M.C.; Ferroni, M.; Sadaoka, Y.; Traversa, E. Screen-printed perovskite-type thick films as gas sensors for environmental monitoring. Sens. Actuators B Chem. 1999, 55, 99–110. [Google Scholar] [CrossRef]
  55. Tanvir, N.B.; Yurchenko, O.; Wilbertz, C.; Urban, G. Investigation of CO2 reaction with copper oxide nanoparticles for room temperature gas sensing. J. Mater. Chem. A 2016, 4, 5294–5302. [Google Scholar] [CrossRef] [Green Version]
  56. Joshi, S.; Ram Kumar, C.B.; Jones, L.A.; Mayes, E.L.H.; Ippolito, S.J.; Sunkara, M.V. Modulating interleaved ZnO assembly with CuO nanoleaves for multifunctional performance: Perdurable CO2 gas sensor and visible light catalyst. Inorg. Chem. Front. 2017, 4, 1848–1861. [Google Scholar] [CrossRef]
  57. Michel, C.R.; Martínez-Preciado, A.H.; Rivera-Tello, C.D. CO2 gas sensing response of YPO4 nanobelts produced by a colloidal method. Sens. Actuators B Chem. 2015, 221, 499–506. [Google Scholar] [CrossRef]
  58. Star, A.; Han, T.-R.; Joshi, V.; Gabriel, J.-C.P.; Grüner, G. Nanoelectronic Carbon Dioxide Sensors. Adv. Mater. 2004, 16, 2049–2052. [Google Scholar] [CrossRef]
  59. Willa, C.; Schmid, A.; Briand, D.; Yuan, J.; Koziej, D. Lightweight, Room-Temperature CO2Gas Sensor Based on Rare-Earth Metal-Free Composites—An Impedance Study. ACS Appl. Mater. Interfaces 2017, 9, 25553–25558. [Google Scholar] [CrossRef] [Green Version]
  60. Klar, T.; Perner, M.; Grosse, S.; Plessen, G.V.; Spirkl, W.; Feldmann, J. Surface-Plasmon Resonances in Single Metallic Nanoparticles. Phys. Rev. Lett. 1998, 80, 4249–4252. [Google Scholar] [CrossRef]
  61. Chan, G.H.; Zhao, J.; Schatz, G.C.; Duyne, R.P.V. Localized Surface Plasmon Resonance Spectroscopy of Triangular Aluminum Nanoparticles. J. Phys. Chem. C 2008, 112, 13958–13963. [Google Scholar] [CrossRef]
  62. Shabaninezhad, M.; Ramakrishna, G. Theoretical Investigation of Plasmonic Properties of Quantum-Sized Silver Nanoparticles. Plasmonics 2019, 15, 783–795. [Google Scholar] [CrossRef]
  63. Leong, H.-S.; Guo, J.; Lindquist, R.G.; Liu, Q.H. Surface plasmon resonance in nanostructured metal films under the Kretschmann configuration. J. Appl. Phys. 2009, 106, 124314. [Google Scholar] [CrossRef] [Green Version]
  64. Jiang, N.; Zhuo, X.; Wang, J. Active Plasmonics: Principles, Structures, and Applications. Chem. Rev. 2017, 118, 3054–3099. [Google Scholar] [CrossRef] [PubMed]
  65. Willets, K.A.; Duyne, R.P.V. Localized Surface Plasmon Resonance Spectroscopy and Sensing. Annu. Rev. Phys. Chem. 2007, 58, 267–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Wang, D.-S.; Fan, S.-K. Microfluidic Surface Plasmon Resonance Sensors: From Principles to Point-of-Care Applications. Sensors 2016, 16, 1175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Jana, J.; Ganguly, M.; Pal, T. Enlightening surface plasmon resonance effect of metal nanoparticles for practical spectroscopic application. RSC Adv. 2016, 6, 86174–86211. [Google Scholar] [CrossRef]
  68. Sadrolhosseini, A.R.; Noor, A.S.M.; Moksin, M.M. Application of Surface Plasmon Resonance Based on Metal Nanoparticle. In Plasmonics—Principles and Applications; InTech: Rijeka, Croatia, 2012. [Google Scholar]
  69. Patil, P.O.; Pandey, G.R.; Patil, A.G.; Borse, V.B.; Deshmukh, P.K.; Patil, D.R.; Tade, R.S.; Nangare, S.N.; Khan, Z.G.; Patil, A.M.; et al. Graphene-based nanocomposites for sensitivity enhancement of surface plasmon resonance sensor for biological and chemical sensing: A review. Biosens. Bioelectron. 2019, 139, 111324. [Google Scholar] [CrossRef] [PubMed]
  70. Rahimi, M.; Singh, J.K.; Babu, D.J.; Schneider, J.J.; Müller-Plathe, F. Understanding Carbon Dioxide Adsorption in Carbon Nanotube Arrays: Molecular Simulation and Adsorption Measurements. J. Phys. Chem. C 2013, 117, 13492–13501. [Google Scholar] [CrossRef]
  71. Cinke, M.; Li, J.; Bauschlicher, C.W.; Ricca, A.; Meyyappan, M. CO2 adsorption in single-walled carbon nanotubes. Chem. Phys. Lett. 2003, 376, 761–766. [Google Scholar] [CrossRef]
  72. Kumar, G.S.; Rao, J.V.R. Detection of NH3 & CO2 using carbon nanotubes at room temperature. Int. J. Nanotechnol. Appl. 2013, 3, 11–18. [Google Scholar]
  73. Nuryadi, R.; Mayasari, R.D. ZnO/Au-based surface plasmon resonance for CO2 gas sensing application. Appl. Phys. A 2015, 122, 33. [Google Scholar] [CrossRef]
  74. Dostálek, J.; Ladd, J.; Jiang, S.; Homola, J. SPR Biosensors for Detection of Biological and Chemical Analytes. In Surface Plasmon Resonance Based Sensors; Springer: Berlin/Heidelberg, Germany, 2006; pp. 177–190. [Google Scholar] [CrossRef]
  75. Olaru, A.; Bala, C.; Jaffrezic-Renault, N.; Aboul-Enein, H.Y. Surface Plasmon Resonance (SPR) Biosensors in Pharmaceutical Analysis. Crit. Rev. Anal. Chem. 2015, 45, 97–105. [Google Scholar] [CrossRef]
  76. Guo, Z.; Song, N.R.; Moon, J.H.; Kim, M.; Jun, E.J.; Choi, J.; Lee, J.Y.; Bielawski, C.W.; Sessler, J.L.; Yoon, J. A Benzobisimidazolium-Based Fluorescent and Colorimetric Chemosensor for CO2. J. Am. Chem. Soc. 2012, 134, 17846–17849. [Google Scholar] [CrossRef] [PubMed]
  77. Ko, K.; Lee, J.-Y.; Chung, H. Highly efficient colorimetric CO2 sensors for monitoring CO2 leakage from carbon capture and storage sites. Sci. Total Environ. 2020, 729, 138786. [Google Scholar] [CrossRef] [PubMed]
  78. Aldewachi, H.; Chalati, T.; Woodroofe, M.N.; Bricklebank, N.; Sharrack, B.; Gardiner, P. Gold nanoparticle-based colorimetric biosensors. Nanoscale 2018, 10, 18–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Rathfon, J.M.; Al-Badri, Z.M.; Shunmugam, R.; Berry, S.M.; Pabba, S.; Keynton, R.S.; Cohn, R.W.; Tew, G.N. Fluorimetric Nerve Gas Sensing Based on Pyrene Imines Incorporated into Films and Sub-Micrometer Fibers. Adv. Funct. Mater. 2009, 19, 689–695. [Google Scholar] [CrossRef]
  80. Chu, C.-S.; Lo, Y.-L. Highly sensitive and linear optical fiber carbon dioxide sensor based on sol–gel matrix doped with silica particles and HPTS. Sens. Actuators B Chem. 2009, 143, 205–210. [Google Scholar] [CrossRef]
  81. Barroso, M.M. Quantum Dots in Cell Biology. J. Histochem. Cytochem. 2011, 59, 237–251. [Google Scholar] [CrossRef] [Green Version]
  82. Hussain, C.M. Handbook of Nanomaterials for Industrial Applications; Elsevier: Amsterdam, The Netherlands, 2018. [Google Scholar]
  83. Ke, J. Semiconductor Nanocrystal–Based Nanosensors and Metal Ions Sensing. Adv. Nanomater. Pollut. Sens. Environ. Catal. 2020, 79–117. [Google Scholar] [CrossRef]
  84. Chung, E.J.; Leon, L.; Rinaldi, C. Nanoparticles for Biomedical Applications: Fundamental Concepts, Biological Interactions and Clinical Applications; Elsevier: Amsterdam, The Netherlands, 2020. [Google Scholar]
  85. Jin, T.; Fujii, F.; Yamada, E.; Nodasaka, Y.; Kinjo, M. Control of the Optical Properties of Quantum Dots by Surface Coating with Calix[n]arene Carboxylic Acids. J. Am. Chem. Soc. 2006, 128, 9288–9289. [Google Scholar] [CrossRef]
  86. Joo, S.-Y.; Park, H.-S.; Kim, D.-Y.; Kim, B.-S.; Lee, C.G.; Kim, W.-B. An investigation into the effective surface passivation of quantum dots by a photo-assisted chemical method. AIP Adv. 2018, 8, 015017. [Google Scholar] [CrossRef] [Green Version]
  87. Saha, S.; Sarkar, P. Controlling the electronic energy levels of ZnO quantum dots using mixed capping ligands. Rsc. Adv. 2014, 4, 1640–1645. [Google Scholar] [CrossRef]
  88. Bera, D.; Qian, L.; Tseng, T.-K.; Holloway, P.H. Quantum Dots and Their Multimodal Applications: A Review. Materials 2010, 3, 2260–2345. [Google Scholar] [CrossRef] [Green Version]
  89. Balogh, K.; Jesus, J.M.; Gouveia, C.; Domingues, J.O.; Markovics, A.; Baptista, J.M.; Kovacs, B.; Pereira, C.M.; Borges, M.-T.; Jorge, P.A.S. Characterization of a novel dissolved CO2sensor for utilization in environmental monitoring and aquaculture industry. In Proceedings of the 8th Iberoamerican Optics Meeting and 11th Latin American Meeting on Optics, Lasers, and Applications, Porto, Portugal, 22–26 July 2013. International Society for Optics and Photonics. [Google Scholar] [CrossRef]
  90. Klantsataya, E.; Jia, P.; Ebendorff-Heidepriem, H.; Monro, T.; François, A. Plasmonic Fiber Optic Refractometric Sensors: From Conventional Architectures to Recent Design Trends. Sensors 2016, 17, 12. [Google Scholar] [CrossRef] [Green Version]
  91. Sharma, A.K.; Jha, R.; Gupta, B.D. Fiber-Optic Sensors Based on Surface Plasmon Resonance: A revie Review. IEEE Sens. J. 2007, 7, 1118–1129. [Google Scholar] [CrossRef]
  92. Kim, K.-J.; Lu, P.; Culp, J.T.; Ohodnicki, P.R. Metal–Organic Framework Thin Film Coated Optical Fiber Sensors: A Novel Waveguide-Based Chemical Sensing Platform. ACS Sens. 2018, 3, 386–394. [Google Scholar] [CrossRef] [PubMed]
  93. Herminjard, S.; Sirigu, L.; Herzig, H.P.; Studemann, E.; Crottini, A.; Pellaux, J.-P.; Gresch, T.; Fischer, M.; Faist, J. Surface Plasmon Resonance sensor showing enhanced sensitivity for CO2 detection in the mid-infrared range. Opt. Express 2009, 17, 293. [Google Scholar] [CrossRef] [PubMed]
  94. Shivananju, B.N.; Yamdagni, S.; Fazuldeen, R.; Kumar, A.K.S.; Hegde, G.M.; Varma, M.M.; Asokan, S. CO2 sensing at room temperature using carbon nanotubes coated core fiber Bragg grating. Rev. Sci. Instrum. 2013, 84, 065002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Chong, X.; Kim, K.-J.; Li, E.; Zhang, Y.; Ohodnicki, P.R.; Chang, C.-H.; Wang, A.X. Near-infrared absorption gas sensing with metal-organic framework on optical fibers. Sens. Actuators B Chem. 2016, 232, 43–51. [Google Scholar] [CrossRef]
  96. Allsop, T.; Arif, R.; Neal, R.; Kalli, K.; Kundrát, V.; Rozhin, A.; Culverhouse, P.; Webb, D.J. Photonic gas sensors exploiting directly the optical properties of hybrid carbon nanotube localized surface plasmon structures. Light Sci. Appl. 2016, 5, e16036. [Google Scholar] [CrossRef] [Green Version]
  97. Njegovec, M.; Donlagic, D. A Fiber-Optic Gas Sensor and Method for the Measurement of Refractive Index Dispersion in NIR. Sensors 2020, 20, 3717. [Google Scholar] [CrossRef]
  98. Wysokiński, K.; Napierała, M.; Stańczykski, T.; Lipiński, S.; Nasiłowski, T. Study on the Sensing Coating of the Optical Fibre CO2 Sensor. Sensors 2015, 15, 31888–31903. [Google Scholar]
  99. Yi, S.; Park, Y.; Han, S.; Min, N.; Kim, E.; Ahn, T. Novel NDIR CO/sub 2/ sensor for indoor air quality monitoring. In Proceedings of the 13th International Conference on Solid-State Sensors, Actuators and Microsystems, 2005. Digest of Technical Papers. TRANSDUCERS ‘05, Seoul, Korea, 5–9 June 2005. [Google Scholar] [CrossRef]
  100. Frodl, R.; Tille, T. A High-Precision NDIR CO2 Gas Sensor for Automotive Applications. IEEE Sens. J. 2006, 6, 1697–1705. [Google Scholar] [CrossRef]
  101. Gibson, D.; Macgregor, C. A Novel Solid State Non-Dispersive Infrared CO2 Gas Sensor Compatible with Wireless and Portable Deployment. Sensors 2013, 13, 7079–7103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Liu, H.; Shi, Y.; Wang, T. Design of a six-gas NDIR gas sensor using an integrated optical gas chamber. Opt. Express 2020, 28, 11451. [Google Scholar] [CrossRef] [PubMed]
  103. International Light Technology. Subminiature T-¾ and T-1 NDIR Lamps. Available online: https://www.intl-lighttech.com/instrumentation-sensor-light-sources/subminiature-t-34-and-t-1-ndir-lamps (accessed on 28 August 2020).
  104. Müller, L.; Käpplinger, I.; Biermann, S.; Brode, W.; Hoffmann, M. Infrared emitting nanostructures for highly efficient microhotplates. J. Micromech. Microeng. 2014, 24, 035014. [Google Scholar] [CrossRef]
  105. Koppens, F.H.L.; Mueller, T.; Avouris, P.; Ferrari, A.C.; Vitiello, M.S.; Polini, M. Photodetectors based on graphene, other two-dimensional materials and hybrid systems. Nat. Nanotechnol. 2014, 9, 780–793. [Google Scholar] [CrossRef]
  106. Pusch, A.; Luca, A.D.; Oh, S.S.; Wuestner, S.; Roschuk, T.; Chen, Y.; Boual, S.; Ali, Z.; Phillips, C.C.; Hong, M.; et al. A highly efficient CMOS nanoplasmonic crystal enhanced slow-wave thermal emitter improves infrared gas-sensing devices. Sci. Rep. 2015, 5, 17451. [Google Scholar] [CrossRef] [PubMed]
  107. Comini, E.; Faglia, G.; Sberveglieri, G. Solid State Gas Sensing; Springer: New York, NY, USA, 2009. [Google Scholar]
  108. Korotcenkov, G. Handbook of Gas Sensor Materials: Properties, Advantages and Shortcomings for Applications; Springer: New York, NY, USA, 2014. [Google Scholar]
  109. Dinh, T.-V.; Choi, I.-Y.; Son, Y.-S.; Kim, J.-C. A review on non-dispersive infrared gas sensors: Improvement of sensor detection limit and interference correction. Sens. Actuators B Chem. 2016, 231, 529–538. [Google Scholar] [CrossRef]
  110. Xu, Q.; Lee, S.; Cho, Y.; Kim, M.H.; Bouffard, J.; Yoon, J. Polydiacetylene-Based Colorimetric and Fluorescent Chemosensor for the Detection of Carbon Dioxide. J. Am. Chem. Soc. 2013, 135, 17751–17754. [Google Scholar] [CrossRef]
  111. Bültzingslöwen, C.V.; Mcevoy, A.K.; Mcdonagh, C.; Maccraith, B.D.; Klimant, I.; Krause, C.; Wolfbeis, O.S. Sol–gel based optical carbon dioxide sensor employing dual luminophore referencing for application in food packaging technology. Analyst 2002, 127, 1478–1483. [Google Scholar] [CrossRef]
  112. Chu, C.-S.; Hsieh, M.-W. Optical fiber carbon dioxide sensor based on colorimetric change of α-naphtholphthalein and CIS/ZnS quantum dots incorporated with a polymer matrix. Opt. Mater. Express 2019, 9, 2937. [Google Scholar] [CrossRef]
  113. Hung, S.-S.; Chang, H.-C.; Chang, I.-N. A Portable Array-Type Optical Fiber Sensing Instrument for Real-Time Gas Detection. Sensors 2016, 16, 2087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Kole, K.; Halder, A.; Singh, S.; Samanta, A.; Das, S.; Kundu, A.K.; Bhattacharyya, D.; Pal, S.K.; Jana, S. Chromogenic-Functionalized Silica Nanoflower Composites for the Detection of Carbon Dioxide. ACS Appl. Nano Mater. 2020, 3, 4321–4328. [Google Scholar] [CrossRef]
  115. Álvarez-Ramos, M.E.; Necochea-Chamorro, J.I.; Carrillo-Torres, R.C.; Sánchez-Zeferino, R. Room temperature CO2 sensing using Au-decorated ZnO nanorods deposited on an optical fiber. Mater. Sci. Eng. B 2020, 262, 114720. [Google Scholar] [CrossRef]
  116. Ali, R.; Saleh, S.M.; Meier, R.J.; Azab, H.A.; Abdelgawad, I.I.; Wolfbeis, O.S. Upconverting nanoparticle based optical sensor for carbon dioxide. Sens. Actuators B Chem. 2010, 150, 126–131. [Google Scholar] [CrossRef]
  117. Chong, X.; Kim, K.-J.; Ohodnicki, P.R.; Li, E.; Chang, C.-H.; Wang, A.X. Ultrashort Near-Infrared Fiber-Optic Sensors for Carbon Dioxide Detection. IEEE Sens. J. 2015, 15, 5327–5332. [Google Scholar] [CrossRef]
  118. Smith, D. CO2 Sensors: Which Type Should You Be Looking For? Available online: https://learn.kaiterra.com/en/air-academy/carbon-dioxide-sensors (accessed on 14 September 2020).
Figure 1. Sensing mechanism of electrochemical CO2 sensors via different nanomaterials.
Figure 1. Sensing mechanism of electrochemical CO2 sensors via different nanomaterials.
Nanomaterials 10 02251 g001
Figure 2. (a) Humidity inhibits sensor’s response (modified from [49]), (b) humidity enhances sensor’s response using LaOCl by two different synthesis techniques sample A (simple oxidation)–sample B (sol-gel) (modified from [48]).
Figure 2. (a) Humidity inhibits sensor’s response (modified from [49]), (b) humidity enhances sensor’s response using LaOCl by two different synthesis techniques sample A (simple oxidation)–sample B (sol-gel) (modified from [48]).
Nanomaterials 10 02251 g002
Figure 3. Mechanism of surface plasmon resonance (SPR) nanomaterial sensing for CO2 (modified from Patil et al., 2019 [69]).
Figure 3. Mechanism of surface plasmon resonance (SPR) nanomaterial sensing for CO2 (modified from Patil et al., 2019 [69]).
Nanomaterials 10 02251 g003
Figure 4. Schematic of colorimetric CO2 sensing. The pH change can be observed on the litmus paper or in the cuvette where the intensity is measured via UV-Vis spectroscopy.
Figure 4. Schematic of colorimetric CO2 sensing. The pH change can be observed on the litmus paper or in the cuvette where the intensity is measured via UV-Vis spectroscopy.
Nanomaterials 10 02251 g004
Figure 5. Quantum dots’ (QDs) response to CO2 sensing based on size, wavelength, and energy.
Figure 5. Quantum dots’ (QDs) response to CO2 sensing based on size, wavelength, and energy.
Nanomaterials 10 02251 g005
Figure 6. The use of nanomaterials thin film in fiber optics.
Figure 6. The use of nanomaterials thin film in fiber optics.
Nanomaterials 10 02251 g006
Figure 7. Schematic of non-dispersive infrared (NDIR) CO2 sensing with potential nanomaterial integration.
Figure 7. Schematic of non-dispersive infrared (NDIR) CO2 sensing with potential nanomaterial integration.
Nanomaterials 10 02251 g007
Table 1. Most significant recent work done in nanomaterial-based electrochemical CO2 sensors.
Table 1. Most significant recent work done in nanomaterial-based electrochemical CO2 sensors.
Nanomaterial
Used
Dynamic Range
(ppm)
Response/Recovery Time (s)MeasurandTemperature (°C)Remarks
CuO nanoparticles [55]400–4000720 @ 0% R.H
500 @ 20% R.H
Surface charge
Surface species
25R.H ↑ – CO2 diff ↓
Bi2O3 nanostructures
[29]
10–100132/82Resistance25S.A/porosity ↑ –adsorption/desorption of CO2
Inorganic silica nanoparticles
[15]
400–2000>60/>150Capacitance30, 42, 58Amine groups ↑–recovery time ↓
p-CuO/n-ZnO hetero-surfaces [56]100076/265Resistance100–400 Optimum = 320lattice mismatch ↑–electron transfer ↑
La1−xSrxFeO3 (O9XS0.3) [47]2000, 4000660/300Resistance200–500
Optimum = 380
R.H ↑– sensitivity ↓
YPO4 nanobelts
[57]
200–800136/N/AImpedance350, 400T > 400 °C–NS ↓
100 nm fumed silica
[36]
500–3000>120Capacitance38–65R.H < 60%, T > 46 °C–response/recovery time, C ↓
200–400 nm SnO nanoparticles
[37]
N/A5/5Resistance25S/V ↑ – response time ↓
LaOCI-doped SnO2 nanofibers [7]100–200024/92Resistance300Porosity ↑ –response/recovery time↓
Doping ↑ 0–8% –response/recovery time ↓
Carbon nanotubes
[58]
500–100,000A few secondsConductance25CO2 ↑ – pH ↓–I ↑
Poly(ionicliquid) alumina
composite [59]
300–3200420/2400Impedance25CO2 ↑ – proton diff ↓
R.H: relative humidity, Diff: diffusion, S.A: surface area, T: temperature, NS: Nanostructure, S/V: surface to volume ratio, I: current, C: capacitance, ↑: increase, ↓: decrease.
Table 2. Most significant recent work undertaken on optical nanosensors.
Table 2. Most significant recent work undertaken on optical nanosensors.
Nanomateiral UsedSensing PrincipleDynamic Range Response Time/Recovery TimeRemarks
polydiacetylene nanofibers [110]Col400 ppminstantaneous/NANaked eye detection for CO2 using green laser pointers.
Ru nano beads doped HPTS in ormosil matrix. [111]Col0–100%30 s/<60 sT ↑ – sensitivity ↓
LOD = 0.08%
CuInS2/ZnS quantum dots. [112]Col0–100%23/71 sT ↑ – sensitivity ↓
Silica nanoparticles porous [113]Ref2–5%48/76 sporosity ↑–sensitivity ↓
CNTs. [94] Ref1000–4000 ppm3.07/2.95 minLOD = 75 ppm
Silica nanoflower [114]Col400–70,000 ppmInstananuous/NACO2 conc. ↑–color intensity ↑
Au-decorated ZnO nanorods [115]Ref0–2000 sccm50/110 sCO2 sensitivity for ZnO–Au lower compared to ZnO
NaYF4:Yb,Er nanoparticles [116]Col0–3%10/180 sLOD = 0.11%
Cu-benzene-1,3, 5-tricarboxylate [117]Ref>500 ppm40/75 sSensing length = 8 cm
Col: colorimetric, Ref: refractometric, ↑: increase, ↓: decrease, LOD: Limit of detection, HPTS: 1-hydroxypyrene-3,6,8-trisulfonate, ormosil: organically modified silica.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rezk, M.Y.; Sharma, J.; Gartia, M.R. Nanomaterial-Based CO2 Sensors. Nanomaterials 2020, 10, 2251. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10112251

AMA Style

Rezk MY, Sharma J, Gartia MR. Nanomaterial-Based CO2 Sensors. Nanomaterials. 2020; 10(11):2251. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10112251

Chicago/Turabian Style

Rezk, Marwan Y., Jyotsna Sharma, and Manas Ranjan Gartia. 2020. "Nanomaterial-Based CO2 Sensors" Nanomaterials 10, no. 11: 2251. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10112251

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