Next Article in Journal
Assessment of PCBs in Surface Waters at Ultratrace Levels: Traditional Approaches and Biomonitoring (Lake Baikal, Russia)
Next Article in Special Issue
Indoor Genetic Algorithm-Based 5G Network Planning Using a Machine Learning Model for Path Loss Estimation
Previous Article in Journal
The Effect of Marketing Activities on Web Search Volume: An Empirical Analysis of Chinese Film Industry Data
Previous Article in Special Issue
Hybrid Network–Spatial Clustering for Optimizing 5G Mobile Networks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Compact and Highly Sensitive Bended Microwave Liquid Sensor Based on a Metamaterial Complementary Split-Ring Resonator

1
Laboratoire d’Electronique de Puissance et Commande Industrielle (LEPCI), Department of Electronics, University of Ferhat Abbas, Sétif -1-, Sétif 19000, Algeria
2
Laboratoire d’Electrotechnique de Skikda (LES), Department of Electrical Engineering, University 20 Aout 1955-Skikda, Skikda 21000, Algeria
3
Instituto de Telecomunicações, Campus Universitário de Santiago, 3810-193 Aveiro, Portugal
4
Faculty of Engineering and Informatics, University of Bradford, Bradford BD7 1DP, UK
5
Department of Electrical and Electronics Engineering, Izmir Katip Celebi University, No:33/2, Izmir 35620, Turkey
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(4), 2144; https://doi.org/10.3390/app12042144
Submission received: 25 December 2021 / Revised: 2 February 2022 / Accepted: 16 February 2022 / Published: 18 February 2022
(This article belongs to the Special Issue 5G Network Planning and Design)

Abstract

:
In this paper, we present the design of a compact and highly sensitive microwave sensor based on a metamaterial complementary split-ring resonator (CSRR), for liquid characterization at microwave frequencies. The design consists of a two-port microstrip-fed rectangular patch resonating structure printed on a 20 × 28 mm2 Roger RO3035 substrate with a thickness of 0.75 mm, a relative permittivity of 3.5, and a loss tangent of 0.0015. A CSRR is etched on the ground plane for the purpose of sensor miniaturization. The investigated liquid sample is put in a capillary glass tube lying parallel to the surface of the sensor. The parallel placement of the liquid test tube makes the design twice as efficient as a normal one in terms of sensitivity and Q factor. By bending the proposed structure, further enhancements of the sensor design can be obtained. These changes result in a shift in the resonant frequency and Q factor of the sensor. Hence, we could improve the sensitivity 10-fold compared to the flat structure. Subsequently, two configurations of sensors were designed and tested using CST simulation software, validated using HFSS simulation software, and compared to structures available in the literature, obtaining good agreement. A prototype of the flat configuration was fabricated and experimentally tested. Simulation results were found to be in good agreement with the experiments. The proposed devices exhibit the advantage of exploring multiple rapid and easy measurements using different test tubes, making the measurement faster, easier, and more cost-effective; therefore, the proposed high-sensitivity sensors are ideal candidates for various sensing applications.

1. Introduction

Wireless sensor networks are essential in our society, and are becoming an integral part of our environment [1,2,3,4,5,6]. These sensors are the object of very active research because of the growing interest in them in many areas, such as environmental monitoring [1,2], industrial control [3], information security [4], the Internet of vehicles [5] and network lifetime [6]. The goal of these studies is to develop sensitive, fast, and easy-to-use sensors. Microwave sensor technology may provide wireless passive sensors in addition to wireless sensor networks. Microwave sensors are the current technological framework that has most attracted the attention of researchers in recent years. These sensors offer many advantages, including low manufacturing costs, high sensitivity, and high durability; these advantages make them very attractive and preferred choices in a variety of research fields, including biomedicine [7,8,9,10], chemistry [11,12,13], electronics, and industry [7,14,15]; they have even recently been used in food safety applications and mechanical systems [16,17,18].
In the field of microwave sensor technology, a new perspective has been developed using the concept of metamaterials [19,20]. These are engineered materials first invented by Veselago [21], composed of subwavelength resonators, and have been experimentally validated in recent years. Metamaterials are of great interest, as they effectively contribute to the design of many new devices, ensuring unusual electromagnetic properties that may not be readily available in nature [22,23], such as the realization of passive double-negative metamaterials (DNMs) for simultaneously negative permittivity (ε) and permeability (μ), and single-negative metamaterials (SNMs), where either (ε) or (μ) is negative, referred to as ENGs and MNGs, respectively [24,25,26,27].
In the earlier literature, a variety of split-ring resonator (SRR) [28,29,30,31] and complementary split-ring resonator (CSRR) [32,33,34,35] sensors have been proposed and studied. SRRs and CSRRs are popular in the design of microwave components, meeting the highest standards of accuracy and sensitivity [36], and show great miniaturization capabilities as they allow a shift of the resonant frequency to lower values [28,37,38]. The disadvantage of SRR-based sensors is that they do not support high electric fields and are not suitable for the detection of large microwave signals. Efforts are constantly deployed to improve microwave sensor systems and overcome this problem. The CSRR-based microwave sensors are based on the electric coupling of CSRRs, which are generally etched in the ground plane and have large electric fields [39], and are the most widely used topology of metamaterials in the design of highly sensitive liquid concentration sensors, due to their low profile and adaptability for various practical applications [40].
In this simulation work, a symmetrical CSRR-based compact and highly sensitive bended microstrip sensor for liquid characterization is presented. This study is based on the flat sensor model studied by Chuma in [41], where the capillary tube placement was arranged to be normal to the surface of the sensor, crossing the patch, the substrate, and the ground plane. Modifications were made to the latter in order to improve its performance. First, the liquid test tube placement was readjusted to lie down on the sensor surface, helping to increase the interaction between the electric field and the examined liquid; this improved both the sensor sensitivity and the Q factor. Second, the flat thin structure was folded around a cylinder of radius R in order to further increase the contact area between the test tube and the sensor surface.
The resulting bended shape strengthens the slow wave propagation within the structure. Thus, the slow wave effect leads to increased interaction time between the electric field and the liquid under test [27]. This phenomenon allows for higher sensitivity. A simulation study was carried out to characterize diverse mixtures of ethanol and water dielectric liquids with different ethanol concentrations. In practice—for example, in engine fuels, pharmaceuticals, and medicinal formulations [41]—ethanol liquids are generally characterized by their high concentration and versatility. The ethanol samples are placed in a glass tube on the surface of the sensor for sensor safety, the possibility of reuse, and ease of operability. This allows rapid analysis of the liquid’s dielectric properties at microwave frequencies.

2. Sensor Structure and Design Steps

The basic flat sensor structure [41], consisting of a two-port microstrip-fed rectangular patch, was printed on a 0.75 mm thick Roger RO3035 substrate with a dielectric constant εr = 3.5 and a loss tangent of tanδ = 0.0015. The structure had overall dimensions of L × W = 20 × 28 mm2 (Figure 1). The ground copper layer held a complementary split-ring resonator (CSRR). The conductors used in this design (patch and ground plane) consisted of 35 µm thick copper layers. The final bended sensor configuration is presented in Figure 2. The novelty of this work lies in the investigation of a new bended-shape structure aiming at enhancing the performance of the microwave sensor. The geometric properties of our proposed sensor are given in Table 1.
The sensitivity of the sensor was analyzed according to the S21 presented results. The shift in the resonant frequency (∆fr = fr_0% − fr_100%) of S21 was deduced. The percentage of relative frequency shift or Q factor (∆fr (3 dB)/fr_ref (%)) was also considered. Next, the sensitivity of the sensors was defined as the ratio of the resonant frequency change to the permittivity change (S = ∆fr/∆εr = |(fr_0% − fr_100%)|/|(εr_0% − εr_100%)|) [42,43].
The S21 parameter and Q factor of the proposed bended sensor, along with CST simulation results of mixing with different water–ethanol concentrations, are presented in Figure 3a,b, respectively. The final designed sensor exhibited a very high sensitivity and an acceptable quality factor compared to those reported in [41]. The frequency range was 400 MHz for a permittivity range of 65 [44] (see Appendix A). This enabled us to achieve a very high sensitivity of 6.15. Moreover, the sensor operates at a central frequency of 1.8 GHz.

3. Simulation, Results, and Discussion

In the following subsections, we present the stepwise evolution of the proposed design, starting from the validation of Chuma’s basic prototype [41], and ending with the final design with the applied modifications—the parallel test tube placement and the bent shape of the sensing structure folded around a cylinder of radius R.

3.1. Flat Structure

Figure 4a illustrates the symmetrical two-port rectangular microstrip-fed patch metamaterial-based flat microwave sensor presented in [41]. A CSRR cell is etched on the ground plane, with a centered hole crossing the patch and substrate, so as to place the test tube of liquid samples normally relative to the surface of the sensor. First, we validated the first basic sensor structure (Figure 4a) by comparing CST simulation results with experiments published in [41], where only experimental results were reported, without simulation validation. The second structure, where the test tube placement was changed to be parallel to the lower side of the sensor from the CSRR side (Figure 4b), was fabricated as illustrated in Figure 5. The simulations showed good agreement with the experiments (Figure 6).
The glass capillary tube was easily produced using a Creality Ender-3 Pro 3D printer using a 0.2 mm nozzle and PLA filament. This was used in the case of the flat structure measurements. This kind of capillary tube is commonly used in clinical and laboratory settings; it is 75 mm long, with an outer radius of 0.75 mm, an inner radius of 0.5 mm, and relative permittivity of 5.5. After the confirmation of the microwave sensing application’s potential through the numerical computations, the microwave sensor model was fabricated on 0.75 mm thick Roger RO3035 substrate with Eleven Lab PCB prototyping system, as shown in Figure 5.
This new tube placement brought a significant benefit in terms of field interaction with the liquid being tested, significantly enhancing the sensor sensitivity.
Figure 6a illustrates the simulated S21 parameter compared to the experimental data reported in [41]. Validation tests on the design show that the simulation results and measurement data are in good agreement for the three considered cases (water concentration 0%, 100%, and air (Figure 6a and Figure 7b)). It should be noted that the higher the concentration of water in the mixture, the lower the frequency of the peak. On the other hand, the resonant frequency and quality factor patterns are nearly overlapping, with a shift of less than 5 MHz. Figure 6b and Figure 7b illustrate the parallel glass tube design simulations of water–ethanol mixture, with an outer radius of r = 0.75 mm. As a result, the sensitivity and quality factor values are doubled compared to the normal test tube case. Note that the normal test tube sensor sensitivity factor was 38 MHz [41], while it was 89 MHz for the parallel case, and the quality factors were 47% and 73%, respectively (Figure 6b and Figure 7b).

3.2. Bended Structure

3.2.1. Effect of the Cylinder’s Radius

In an effort to obtain the highest possible sensitivity factor, various modifications were made to the original configuration. The flat structure was bent into a cylindrical form, and the original size remained unchanged (Figure 8a). Simulations were performed for different values of the cylinder radius R. The results for a test tube of radius R = 3.75 mm are presented in Figure 8b. The response S21 was affected by the bended form of the sensing structure. The frequency band increased proportionally with smaller bending radii. It can be seen that the frequency interval between the upper (100% water) and lower (0% water) water concentrations increases with the decrease in the bending radius for 10 mm, 7 mm, and the minimum possible value of 5 mm. Beyond that, the structure is no longer practically feasible, due to structure deformation; consequently, the sensitivity increases, and reaches its maximum.
Bending the structure on a smaller cylinder radius improves and increases the sensor’s sensitivity. Table 2 summarizes the main obtained results.

3.2.2. Effect of the Test Tube Radius

In addition to its quick and easy handling, the test tube’s parallel position configuration enabled us to investigate more possible structure cases according to the tube radius, for optimization purposes. However, the normal configuration [41] allowed only a single radius case (r = 0.75 mm), since the tube was inserted through a hole drilled in the substrate.
Figure 9 presents a CST simulation validated by Ansys HFSS for a test tube of radius r = 0.75 mm. The two software S21 response results at resonance were similar for different concentrations, with a 50 MHz offset (Figure 9a,b). A sensor sensitivity of 1.7 kHz and a quality factor of 88.35% were achieved using CST. These results were close to those obtained at HFSS 1.5 kHz and 62.25%, respectively. For this case, the device operates at around 2.4 GHz.
Similarly, as shown in Figure 10, the sensitivity increased almost 2.5-fold compared to the case with r = 0.75 mm, while the quality factor (which has a direct link with the imaginary component of the permittivity) decreased by 15%. The CST and HFSS results curves almost overlap. The resonant frequency of this case decreased, averaging 2.2 GHz.
For r = 2.5 mm, an improvement in the performance of the proposed sensor was noticed. The frequency band decreased and reached an average value of 2 GHz at resonance, with stable sensitivity and a slight decrease in the quality factor. In addition, the CST- and HFSS-simulated S21 results are quite close—almost superimposed (Figure 11).
For r = 3.75 mm (Figure 12), the simulation results were closer, and almost overlapped at lower concentrations. On the other hand, they moved slightly further apart at other concentrations, and a remarkable increase in sensitivity compared to the previous cases was observed. In the same way, a slight decrease in the quality factor was observed. The resonant frequency decreased until it reached 1.8 GHz.
In the same way, for r = 5 mm (Figure 13), the sensitivity was improved by more than 3 times relative to that of r = 0.75 mm, with a decrease in the quality factor by 30%. The S21 HFSS validation results were close to those obtained by CST for different concentrations of water in the mixture. The resonant frequency reached values of around 1.4 GHz.
According to Table 3, the accuracy and reliability of the proposed device was improved through five measurements for different tube radii. For instance, tubes with r = 0.75 and 1.5 mm had higher Q factors, and were mainly used to extract the imaginary part (ε″). Tubes with larger radii (r = 3.75 and 5 mm) had a sensitivity range frequency of 400 MHz; hence, they were advantageously used for permittivity and the real part (ε′) extraction, since ε′ and ε″ are directly related to Δfr and to Q factor, respectively [45].

3.3. Complex Permittivity Extraction

The least squares method [32] was used to determine the complex permittivity ( ε x + j ε x ) for different concentrations of ethanol in water. The complex permittivity functions are related to the change in the resonant frequency and the Q factor. The change in the resonant frequency and the Q factor is described in terms of the complex permittivity of the liquid sample, using the linear equations [32,41] translated into matrix M (Equation (1)). For a mathematical analysis allowing the determination of the matrix coefficients, the reference values of the complex permittivity for different concentrations were extracted from [43]. These values are given in tables and illustrated by figures in Appendix A, so as to provide a better understanding of water–ethanol mixture concentrations for important frequency ranges.
[ Δ f Δ Q ] = [ m 11 m 12 m 21 m 22 ] [ Δ ε Δ ε ]
where:
Δ f = Δ f s a m p l e Δ f r e f e r e n c e
Δ Q = Δ Q s a m p l e Δ Q r e f e r e n c e
Δ ε = Δ ε s a m p l e Δ ε r e f e r e n c e
where m11, m12, m21, and m22 are unknown coefficients, Δ ε is the complex permittivity, Δ f is the resonant frequency, and Δ Q is the Q factor. The reference value for analysis purposes is taken at 50% water concentration. The shifts in the values related to the variation in the complex permittivity, the resonant frequency, and the Q factor for different samples of the water–ethanol mixture are defined by the matrices X, Y1, and Y2, respectively, as follows:
X = [ Δ ε 0 % Δ ε 0 % Δ ε 10 % Δ ε 10 % Δ ε 20 % Δ ε 20 % Δ ε 30 % Δ ε 30 % Δ ε 40 % Δ ε 40 % Δ ε 50 % Δ ε 50 % Δ ε 60 % Δ ε 60 % Δ ε 70 % Δ ε 70 % Δ ε 80 % Δ ε 80 % Δ ε 90 % Δ ε 90 % Δ ε 100 % Δ ε 100 % ] = [ 34.2259 5.2866 26.6591 1.9484 20.3808 0.651 12.1428 0.4194 7.7491 0.3072 0 0 5.5436 0.1577 16.0126 0.526 21.5426 2.4037 28.4697 5.009 33.4518 6.732 ] ,   Y 1 = [ Δ f 0 % Δ f 10 % Δ f 20 % Δ f 30 % Δ f 40 % Δ f 50 % Δ f 60 % Δ f 70 % Δ f 80 % Δ f 90 % Δ f 100 % ] = [ 0.3124 0.1867 0.1189 0.0514 0.0365 0 0.0212 0.0501 0.0626 0.0742 0.082 ]   and   Y 2 = [ Δ Q 0 % Δ Q 10 % Δ Q 20 % Δ Q 30 % Δ Q 40 % Δ Q 50 % Δ Q 60 % Δ Q 70 % Δ Q 80 % Δ Q 90 % Δ Q 100 % ] = [ 4.5466 4.7513 4.0621 2.7325 1.6591 0 1.4484 4.0915 7.3011 13.3947 18.0188 ]
After having obtained the complex permittivity, the resonant frequency, and the Q factor for the different samples of the water–ethanol mixtures, we had only to apply the initial matrix model given in (1). Thanks to a detailed algebraic analysis, Equations (5) and (6) were generated [32]. Thus, the unknown coefficients can be determined as given in (7):
[ m 11 m 12 ] T = ( X T X ) 1 X T Y 1
[ m 11 m 12 ] T = ( X T X ) 1 X T Y 1
The final matrix providing data on the different water–ethanol samples is described by (8). These calculated coefficients are used to determine the complex permittivity of any given sample as a function of the change in the sample’s resonant frequency and Q factor:
[ m 11 m 12 m 21 m 22 ] = [ 0.0057 0.0187 0.2734 1.1230 ]
[ Δ ε Δ ε ] = [ m 11 m 12 m 21 m 22 ] 1 [ Δ f r e s Δ Q ] = [ 97.4829 1.6231 23.7329 0.4953 ] [ Δ f r e s Δ Q ]
Table 4 shows a comparison of the proposed sensing structures with the literature in terms of sensitivity and Q factor.

4. Conclusions

This paper proposes two small, low-cost, and highly sensitive microwave CSRR-based sensors for liquid characterization. The capillary glass test tube is filled with a water–ethanol mixture placed parallel to the sensor’s ground surface in order to determine the dielectric parameters of different mixture concentrations. The CSRR-based bended sensor was designed by folding the flat sensor around an arbitrary cylinder. Very important results were obtained. The originality of the bended sensor offers the possibility of changing the test tube quickly and easily for multiple measurements. The diversity of the sensor’s bended shape and the modification of the radius of the test tube significantly improve the sensitivity at different resonant frequencies (1.8 GHz, 1.9 GHz, 2 GHz, and 2.4 GHz). The proposed sensors have several advantages, such as small size, low cost, multi-performance ability, high sensitivity, and easy handling, making them ideal candidates for various applications.

Author Contributions

Conceptualization, S.M. (Said Mosbah), C.Z. and D.S.; methodology, S.M. (Said Mosbah), I.E. and M.P.; software, S.M. (Said Mosbah), M.L.B., S.M. (Samira Mekki) and R.Z.; validation, I.E., C.Z. and D.S.; formal analysis, S.M. (Samira Mekki); investigation, M.P. and J.R.; writing—original draft preparation, S.M. (Said Mosbah), C.Z. and D.S.; writing—review and editing, M.L.B., S.M. (Samira Mekki) and I.E.; visualization, S.M. (Said Mosbah); supervision, R.A.A.-A. and C.Z.; project administration, J.R. and I.E.; funding acquisition, J.R. and I.E. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Moore4Medical project, funded within ECSEL JU in collaboration with the EU H2020 Framework Programme (H2020/2014–2020) under grant agreement H2020-ECSEL-2019-IA-876190, and the Fundação para a Ciência e Tecnologia (ECSEL/0006/2019). This project received funding in part from the DGRSDT (Direction Générale de la Recherche Scientifique et du Développement Technologique), MESRS (Ministry of Higher Education and Scientific Research), Algeria.

Data Availability Statement

All data are included within the manuscript.

Acknowledgments

This work was also supported by the General Directorate of Scientific Research and Technological Development (DGRSDT)–Ministry of Higher Education and Scientific Research (MESRS), Algeria, and funded by the FCT/MEC through national funds and, when applicable, co-financed by the ERDF, under the PT2020 Partnership Agreement under the UID/EEA/50008/2020 project.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

The properties of water–ethanol mixture concentration and complex permittivity are given in [44], and are represented in the following figures. The difficulty encountered when reading the data as described in [44] forced us to reproduce these data in Figure A1. Two tables describing the properties of the water–ethanol (real and imaginary parts of permittivity) mixture are given in Table A1 and Table A2 from [44]. In this appendix, the work was based mainly on the work presented in [44], so as to provide the frequency dependence of the complex permittivity of any water concentration (Figure A1), which will serve as a basis for future research. In our case, the work is limited to the 1.5–3.5 GHz frequency band for water–ethanol mixtures (Table A1 and Table A2).
Figure A1. Frequency dependence of complex permittivity [44]: (a) real part; (b) imaginary part.
Figure A1. Frequency dependence of complex permittivity [44]: (a) real part; (b) imaginary part.
Applsci 12 02144 g0a1
Table A1. Measured real dielectric spectra for water–ethanol mixture at 25 °C and different volume fractions; data reported from [44].
Table A1. Measured real dielectric spectra for water–ethanol mixture at 25 °C and different volume fractions; data reported from [44].
F(GHz)0%10%20%30%40%50%60%70%80%90%100%
1.512.851921.086727.56635.224739.652247.019252.253562.376967.388973.624878.4483
1.612.33520.43326.88234.65739.06746.53651.83062.00967.13673.5278.379
1.711.872119.840926.238834.146138.512846.073151.429661.661766.909273.435778.3121
1.811.323419.17925.557233.575337.980645.613151.030361.33866.686873.339278.2389
1.910.854118.574324.910233.039837.447245.142150.621161.007166.44773.233278.1716
210.424417.991224.269532.507536.901244.650350.193960.662966.192973.1278.1021
2.110.018817.418323.627431.964736.34144.138649.749960.30465.930773.003878.0283
2.29.642716.867922.99531.424435.773843.611249.294659.934465.667572.889777.9499
2.39.30416.354922.386130.905435.208543.074548.833959.559865.411172.781877.8677
2.49.000615.879621.805130.410434.649342.538948.372259.183765.167672.680677.7838
2.58.726515.433821.250229.931934.099242.013447.912158.80864.936172.582877.7011
2.68.477115.011420.719429.46333.561541.497747.451758.431864.705672.480477.6186
2.78.248714.609820.211228.996833.036240.982646.985158.051564.460772.361577.5328
2.88.039414.22819.72528.526532.519340.462646.509957.665264.192772.217877.4422
2.97.848913.864919.261328.04932.008739.942446.030457.274863.90372.050377.3496
37.675413.51818.81927.565931.502939.426145.548756.87963.594571.86377.2556
Table A2. Measured imaginary dielectric spectra for water–ethanol mixture at 25 °C and different volume fractions; data reported from [44].
Table A2. Measured imaginary dielectric spectra for water–ethanol mixture at 25 °C and different volume fractions; data reported from [44].
F(GHz)0%10%20%30%40%50%60%70%80%90%100%
1.59.85712.079112.483612.532911.556611.334211.066610.46248.85257.05445.5275
1.69.63912.097712.708612.964712.020211.902311.658411.10579.41867.47215.9109
1.79.42212.086912.884713.344512.459612.467612.245711.743610.00727.886.2745
1.89.241312.147413.116813.772112.927613.032312.833612.369110.58788.29446.6377
1.99.048612.180613.314614.174613.383713.587713.409512.989211.16878.70797.0145
28.844512.182713.480114.550513.823914.131113.973413.605111.72749.12217.3991
2.18.63412.150713.61114.894714.241114.657314.522414.215712.25639.53727.7857
2.28.419812.088213.705615.202914.629915.156615.053814.81812.76679.95278.1704
2.38.203712.002913.765915.474114.988715.62215.569515.407913.27710.36888.5512
2.47.991411.901113.797315.708815.317416.05716.07315.982213.791510.78688.9274
2.57.788511.787313.806615.908915.617716.470816.566116.539314.300311.20939.2995
2.67.594911.666413.800216.080115.895216.868817.050917.081114.798911.63999.6675
2.77.40811.542613.782316.230416.156817.251217.530817.610915.293912.081910.0337
2.87.229211.417113.75316.363516.403517.61518.004218.128115.788512.533210.4033
2.97.061711.28913.709216.478916.630317.954318.461618.626916.273212.985810.78
36.902911.157413.650416.576816.833518.264218.894719.105316.736713.434611.1603

References

  1. Gao, M.; Pan, J.-S.; Li, J.; Zhang, Z.; Chai, Q.-W. 3-D Terrains Deployment of Wireless Sensors Network by Utilizing Parallel Gases Brownian Motion Optimization. J. Internet Technol. 2021, 22, 13–29. [Google Scholar]
  2. Zegadi, R.; Lorrain, N.; Bodiou, L.; Guendouz, M.; Ziet, L.; Charrier, J. Enhanced Mid-Infrared Gas Absorption Spectroscopic Detection Using Chalcogenide or Porous Germanium Waveguides. J. Opt. 2021, 23, 035102. [Google Scholar] [CrossRef]
  3. Zegadi, R.; Ziet, L.; Zegadi, A. Design of High Sensitive Temperature Sensor Based on Two-Dimensional Photonic Crystal. Silicon 2020, 12, 2133–2139. [Google Scholar] [CrossRef]
  4. Yuan, C.; Sun, X. Fingerprint Liveness Detection Adapted to Different Fingerprint Sensors Based on Multiscale Wavelet Transform and Rotation-Invarient Local Binary Pattern. J. Internet Technol. 2018, 19, 91–98. [Google Scholar]
  5. Malekian, R.; Thakur, A.; Nair, L.; Pedersen, C.F. A Sensor Based Peer to Peer Vehicle Data Sharing System, an Internet of Vehicles Approach. J. Internet Technol. 2018, 19, 2155–2162. [Google Scholar]
  6. Kim, K.T.; Youn, H.Y. A Dynamic Level-Based Routing Protocol for Energy Efficiency in Wireless Sensor Networks. J. Internet Technol. 2017, 18, 11–21. [Google Scholar]
  7. Chudpooti, N.; Duangrit, N.; Sangpet, P.; Akkaraekthalin, P.; Imberg, B.U.; Robertson, I.D.; Somjit, N. In-Situ Self-Aligned NaCl-Solution Fluidic-Integrated Microwave Sensors for Industrial and Biomedical Applications. IEEE Access 2020, 8, 188897–188907. [Google Scholar] [CrossRef]
  8. Entesari, K.; Helmy, A.A.; Moslehi-Bajestan, M. Integrated Systems for Biomedical Applications: Silicon-Based RF\/Microwave Dielectric Spectroscopy and Sensing. IEEE Microw. Mag. 2017, 18, 57–72. [Google Scholar] [CrossRef]
  9. Vlachogiannakis, G.; Hu, Z.; Shivamurthy, H.T.; Neto, A.; Pertijs, M.A.P.; de Vreede, L.C.N.; Spirito, M. Miniaturized Broadband Microwave Permittivity Sensing for Biomedical Applications. IEEE J. Electromagn. RF Microw. Med. Biol. 2019, 3, 48–55. [Google Scholar] [CrossRef]
  10. Bourqui, J.; Fear, E.C. Shielded UWB Sensor for Biomedical Applications. IEEE Antennas Wirel. Propag. Lett. 2012, 11, 1614–1617. [Google Scholar] [CrossRef]
  11. Helmy, A.A.; Jeon, H.-J.; Lo, Y.-C.; Larsson, A.J.; Kulkarni, R.; Kim, J.; Silva-Martinez, J.; Entesari, K. A Self-Sustained CMOS Microwave Chemical Sensor Using a Frequency Synthesizer. IEEE J. Solid State Circuits 2012, 47, 2467–2483. [Google Scholar] [CrossRef]
  12. Mirzaei, A.; Neri, G. Microwave-Assisted Synthesis of Metal Oxide Nanostructures for Gas Sensing Application: A Review. Sens. Actuators B Chem. 2016, 237, 749–775. [Google Scholar] [CrossRef]
  13. Zarifi, M.H.; Deif, S.; Abdolrazzaghi, M.; Chen, B.; Ramsawak, D.; Amyotte, M.; Vahabisani, N.; Hashisho, Z.; Chen, W.; Daneshmand, M. A Microwave Ring Resonator Sensor for Early Detection of Breaches in Pipeline Coatings. IEEE Trans. Ind. Electron. 2018, 65, 1626–1635. [Google Scholar] [CrossRef]
  14. Jilani, M.T.; Rehman, M.Z.U.; Khan, A.M.; Chughtai, O.; Abbas, M.A.; Khan, M.T. An Implementation of IoT-Based Microwave Sensing System for the Evaluation of Tissues Moisture. Microelectron. J. 2019, 88, 117–127. [Google Scholar] [CrossRef]
  15. Bakır, M.; Karaaslan, M.; Unal, E.; Akgol, O.; Sabah, C. Microwave Metamaterial Absorber for Sensing Applications. Opto Electron. Rev. 2017, 25, 318–325. [Google Scholar] [CrossRef]
  16. Kurmendra; Kumar, R. A Review on RF Micro-Electro-Mechanical-Systems (MEMS) Switch for Radio Frequency Applications. Microsyst. Technol. 2021, 27, 2525–2542. [Google Scholar] [CrossRef]
  17. Zhang, K.; Bariani, F.; Dong, Y.; Zhang, W.; Meystre, P. Proposal for an Optomechanical Microwave Sensor at the Subphoton Level. Phys. Rev. Lett. 2015, 114, 113601. [Google Scholar] [CrossRef] [Green Version]
  18. Rawat, V.; Dhobale, S.; Kale, S.N. Ultra-Fast Selective Sensing of Ethanol and Petrol Using Microwave-Range Metamaterial Complementary Split-Ring Resonators. J. Appl. Phys. 2014, 116, 164106. [Google Scholar] [CrossRef]
  19. Petrin, A. Wave Propagation; IntechOpen: London, UK, 2011; ISBN 978-953-307-275-3. [Google Scholar]
  20. Ebrahimi, A.; Withayachumnankul, W.; Al-Sarawi, S.; Abbott, D. High-Sensitivity Metamaterial-Inspired Sensor for Microfluidic Dielectric Characterization. IEEE Sens. J. 2014, 14, 1345–1351. [Google Scholar] [CrossRef] [Green Version]
  21. Veselago, V.G. The electrodynamics of substances with simultaneously negative values of ε and μ. Sov. Phys. Usp. 1968, 10, 509–514. [Google Scholar] [CrossRef]
  22. Sarkhel, A.; Mitra, D.; Chaudhuri, S.R.B. A Compact Metamaterial with Multi-Band Negative-Index Characteristics. Appl. Phys. A 2016, 122, 471. [Google Scholar] [CrossRef]
  23. Alqadami, A.S.M.; Jamlos, M.F.; Soh, P.J.; Rahim, S.K.A.; Vandenbosch, G.A.E.; Narbudowicz, A. Miniaturized Dual-Band Antenna Array with Double-Negative (DNG) Metamaterial for Wireless Applications. Appl. Phys. A 2016, 123, 22. [Google Scholar] [CrossRef]
  24. Webb, B.A.; Ziolkowski, R.W. Metamaterial-Inspired Multilayered Structures Optimized to Enable Wireless Communications through a Plasmasonic Region. Appl. Phys. Lett. 2021, 118, 094102. [Google Scholar] [CrossRef]
  25. Manage, P.S.; Naik, U.; Nargundkar, S.; Rayar, V. A Survey on Applications of Metamaterials in Antenna Design. In Proceedings of the 2020 Third International Conference on Smart Systems and Inventive Technology (ICSSIT), Tirunelveli, India, 20–22 August 2020; pp. 153–158. [Google Scholar]
  26. Canet-Ferrer, J. Metamaterials and Metasurfaces; IntechOpen: London, UK, 2019; ISBN 978-1-78984-842-7. [Google Scholar]
  27. Kayal, S.; Shaw, T.; Mitra, D. Design of Metamaterial-Based Compact and Highly Sensitive Microwave Liquid Sensor. Appl. Phys. A 2019, 126, 13. [Google Scholar] [CrossRef]
  28. Muhammed, S.K.T.; Ansari, M.A.H.; Jha, A.K.; Akhtar, M.J. Design of SRR-Based Microwave Sensor for Characterization of Magnetodielectric Substrates. IEEE Microw. Wirel. Compon. Lett. 2017, 27, 524–526. [Google Scholar] [CrossRef]
  29. Kiani, S.; Rezaei, P.; Navaei, M. Dual-Sensing and Dual-Frequency Microwave SRR Sensor for Liquid Samples Permittivity Detection. Measurement 2020, 160, 107805. [Google Scholar] [CrossRef]
  30. Viswanathan, A.P.; Moolat, R.; Mani, M.; Va, S.; Pezholil, M. A Simple Electrically Small Microwave Sensor Based on Complementary Asymmetric Single Split Resonator for Dielectric Characterization of Solids and Liquids. Int. J. RF Microw. Comput. Aided Eng. 2020, 30, e22462. [Google Scholar] [CrossRef]
  31. Vélez, P.; Su, L.; Grenier, K.; Mata-Contreras, J.; Dubuc, D.; Martín, F. Microwave Microfluidic Sensor Based on a Microstrip Splitter/Combiner Configuration and Split Ring Resonators (SRRs) for Dielectric Characterization of Liquids. IEEE Sens. J. 2017, 17, 6589–6598. [Google Scholar] [CrossRef] [Green Version]
  32. Javed, A.; Arif, A.; Zubair, M.; Mehmood, M.Q.; Riaz, K. A Low-Cost Multiple Complementary Split-Ring Resonator-Based Microwave Sensor for Contactless Dielectric Characterization of Liquids. IEEE Sens. J. 2020, 20, 11326–11334. [Google Scholar] [CrossRef]
  33. Zhao, W.-S.; Gan, H.-Y.; He, L.; Liu, Q.; Wang, D.-W.; Xu, K.; Chen, S.; Dong, L.; Wang, G. Microwave Planar Sensors for Fully Characterizing Magneto-Dielectric Materials. IEEE Access 2020, 8, 41985–41999. [Google Scholar] [CrossRef]
  34. Gan, H.-Y.; Zhao, W.-S.; Liu, Q.; Wang, D.-W.; Dong, L.; Wang, G.; Yin, W.-Y. Differential Microwave Microfluidic Sensor Based on Microstrip Complementary Split-Ring Resonator (MCSRR) Structure. IEEE Sens. J. 2020, 20, 5876–5884. [Google Scholar] [CrossRef]
  35. Lee, C.-S.; Bai, B.; Song, Q.-R.; Wang, Z.-Q.; Li, G.-F. Open Complementary Split-Ring Resonator Sensor for Dropping-Based Liquid Dielectric Characterization. IEEE Sens. J. 2019, 19, 11880–11890. [Google Scholar] [CrossRef]
  36. Armghan, A.; Alanazi, T.M.; Altaf, A.; Haq, T. Characterization of Dielectric Substrates Using Dual Band Microwave Sensor. IEEE Access 2021, 9, 62779–62787. [Google Scholar] [CrossRef]
  37. Kumar, A.; Rajawat, M.S.; Mahto, S.K.; Sinha, R. Metamaterial-Inspired Complementary Split Ring Resonator Sensor and Second-Order Approximation for Dielectric Characterization of Fluid. J. Electron. Mater. 2021, 50, 5925–5932. [Google Scholar] [CrossRef]
  38. Zarifi, M.H.; Thundat, T.; Daneshmand, M. High Resolution Microwave Microstrip Resonator for Sensing Applications. Sens. Actuators A Phys. 2015, 233, 224–230. [Google Scholar] [CrossRef]
  39. Haq, T.; Ruan, C.; Ullah, S.; Kosar, A. Reconfigurable Ultra Wide Band Notch Filter Based on Complementary Metamaterial. In Proceedings of the 2018 IEEE Asia-Pacific Conference on Antennas and Propagation (APCAP), Auckland, New Zealand, 5–8 August 2018; pp. 381–382. [Google Scholar]
  40. Haq, T.; Ruan, C.; Zhang, X.; Ullah, S. Complementary Metamaterial Sensor for Nondestructive Evaluation of Dielectric Substrates. Sensors 2019, 19, 2100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Chuma, E.L.; Iano, Y.; Fontgalland, G.; Bravo Roger, L.L. Microwave Sensor for Liquid Dielectric Characterization Based on Metamaterial Complementary Split Ring Resonator. IEEE Sens. J. 2018, 18, 9978–9983. [Google Scholar] [CrossRef]
  42. Yeo, J.; Lee, J.-I. High-Sensitivity Microwave Sensor Based on an Interdigital-Capacitor-Shaped Defected Ground Structure for Permittivity Characterization. Sensors 2019, 19, 498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Salim, A.; Lim, S. Complementary Split-Ring Resonator-Loaded Microfluidic Ethanol Chemical Sensor. Sensors 2016, 16, 1802. [Google Scholar] [CrossRef] [Green Version]
  44. Bao, J.; Swicord, M.L.; Davis, C.C. Microwave Dielectric Characterization of Binary Mixtures of Water, Methanol, and Ethanol. J. Chem. Phys. 1996, 104, 4441–4450. [Google Scholar] [CrossRef]
  45. Li, D.; Free, C.E.; Pitt, K.E.G.; Barnwell, P.G. A Simple Method for Accurate Loss Tangent Measurement of Dielectrics Using a Microwave Resonant Cavity. IEEE Microw. Wirel. Compon. Lett. 2001, 11, 118–120. [Google Scholar] [CrossRef]
  46. Withayachumnankul, W.; Jaruwongrungsee, K.; Tuantranont, A.; Fumeaux, C.; Abbott, D. Metamaterial-Based Microfluidic Sensor for Dielectric Characterization. Sens. Actuators A Phys. 2013, 189, 233–237. [Google Scholar] [CrossRef] [Green Version]
  47. Awang, R.A.; Tovar-Lopez, F.J.; Baum, T.; Sriram, S.; Rowe, W.S.T. Meta-Atom Microfluidic Sensor for Measurement of Dielectric Properties of Liquids. J. Appl. Phys. 2017, 121, 094506. [Google Scholar] [CrossRef]
Figure 1. Geometric dimensions of the flat sensor structure [41]: (a) side view; (b) top view; (c) bottom view.
Figure 1. Geometric dimensions of the flat sensor structure [41]: (a) side view; (b) top view; (c) bottom view.
Applsci 12 02144 g001
Figure 2. 3D view of the proposed sensor: (a) upper side view; (b) inner side view; (c) final structure with embedded glass tube.
Figure 2. 3D view of the proposed sensor: (a) upper side view; (b) inner side view; (c) final structure with embedded glass tube.
Applsci 12 02144 g002
Figure 3. CST simulations of the flat sensor: (a) S21 parameter and (b) Q factor of mixtures with different water–ethanol concentrations for the proposed sensor, with test glass radius r = 3.75 mm.
Figure 3. CST simulations of the flat sensor: (a) S21 parameter and (b) Q factor of mixtures with different water–ethanol concentrations for the proposed sensor, with test glass radius r = 3.75 mm.
Applsci 12 02144 g003
Figure 4. (a) Sensor structure with a normal glass tube [41]; (b) flat sensor with a parallel glass tube.
Figure 4. (a) Sensor structure with a normal glass tube [41]; (b) flat sensor with a parallel glass tube.
Applsci 12 02144 g004
Figure 5. Fabricated sensor prototype: (a) top view; (b) bottom view; (c) flat sensor with embedded tube.
Figure 5. Fabricated sensor prototype: (a) top view; (b) bottom view; (c) flat sensor with embedded tube.
Applsci 12 02144 g005
Figure 6. S21 parameter simulations of (a) water, ethanol, and air samples of normal glass tube design compared to measurements [41]; (b) water–ethanol concentration of the designed parallel test tube compared to measurements.
Figure 6. S21 parameter simulations of (a) water, ethanol, and air samples of normal glass tube design compared to measurements [41]; (b) water–ethanol concentration of the designed parallel test tube compared to measurements.
Applsci 12 02144 g006
Figure 7. Q factor and resonant frequency versus water–ethanol concentration of (a) normal glass tube design compared to measurements [41], and (b) the designed parallel test tube compared to measurements.
Figure 7. Q factor and resonant frequency versus water–ethanol concentration of (a) normal glass tube design compared to measurements [41], and (b) the designed parallel test tube compared to measurements.
Applsci 12 02144 g007
Figure 8. Designed sensor in the CST interface: (a) bended structure; (b) S21 response of water–ethanol mixtures for different bending radii and concentrations.
Figure 8. Designed sensor in the CST interface: (a) bended structure; (b) S21 response of water–ethanol mixtures for different bending radii and concentrations.
Applsci 12 02144 g008
Figure 9. Simulated results: (a) S21 response and (b) resonant frequency and Q factor for different mixture water concentrations, with r = 0.75 mm.
Figure 9. Simulated results: (a) S21 response and (b) resonant frequency and Q factor for different mixture water concentrations, with r = 0.75 mm.
Applsci 12 02144 g009
Figure 10. Simulated results: (a) S21 response and (b) resonant frequency and Q factor for different mixture water concentrations, with r = 1.5 mm.
Figure 10. Simulated results: (a) S21 response and (b) resonant frequency and Q factor for different mixture water concentrations, with r = 1.5 mm.
Applsci 12 02144 g010
Figure 11. Simulated results: (a) S21 response and (b) resonant frequency and Q factor for different mixture water concentrations, with r = 2.5 mm.
Figure 11. Simulated results: (a) S21 response and (b) resonant frequency and Q factor for different mixture water concentrations, with r = 2.5 mm.
Applsci 12 02144 g011
Figure 12. Simulated results: (a) S21 response and (b) resonant frequency and Q factor for different mixture water concentrations, with r = 3.75 mm.
Figure 12. Simulated results: (a) S21 response and (b) resonant frequency and Q factor for different mixture water concentrations, with r = 3.75 mm.
Applsci 12 02144 g012
Figure 13. Simulated results: (a) S21 response and (b) resonant frequency and Q factor for different mixture water concentrations, with r = 5 mm.
Figure 13. Simulated results: (a) S21 response and (b) resonant frequency and Q factor for different mixture water concentrations, with r = 5 mm.
Applsci 12 02144 g013
Table 1. Geometrical parameters of the proposed sensor structure.
Table 1. Geometrical parameters of the proposed sensor structure.
ParametersWLHsW1W2W3W4L1L2L3L4L5
Value (mm)28200.756.23.62.941.69.28.546.21.62.26
ParametersL6d1d2d3d4d5d6a1a2a3sθ
Value (mm)0.946.25.023.842.852.261.670.330.330.330.545°
Table 2. Effect of bending radius on the sensitivity Q factor of the bent sensor.
Table 2. Effect of bending radius on the sensitivity Q factor of the bent sensor.
Bending RadiusResonant Frequency 0% (MHz)Resonant Frequency 100% (MHz)Δfr
(MHz)
Maximum Q Factor
R = 10 mm2295.62046.3249.338.6
R = 7.5 mm2225.61936.4289.234
R = 5 mm1990.71596.3394.527.9
Table 3. Sensitivity for different tube radius.
Table 3. Sensitivity for different tube radius.
Tube Radius
(mm)
Resonant Frequency 0% (MHz)Resonant Frequency 100% (MHz)Δfr
(MHz)
Maximum Q Factor
(%)
0.752403.72284.3119.488.4
1.52303.62047.8255.853.2
2.52061.11819.6241.544.1
3.751990.71596.3394.527.9
51758.61376.9381.723.3
Table 4. Comparison of the proposed sensor with data reported in the literature.
Table 4. Comparison of the proposed sensor with data reported in the literature.
Reference0% Water Concentration Resonant Frequency (MHz)100% Water Concentration Resonant Frequency (MHz)Δf (MHz)Maximum Q Factor
[18]398042502707
[20]192015303909
[32]2370202035032
[34]105015004505
[35]265210558
[37]199719623542
[41]234823024647
[46]1960185510526
[47]305029906055
Proposed flat structure2418.52330.488.172.71
Proposed bended structure1990.71596.3394.527.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mosbah, S.; Zebiri, C.; Sayad, D.; Elfergani, I.; Bouknia, M.L.; Mekki, S.; Zegadi, R.; Palandoken, M.; Rodriguez, J.; Abd-Alhameed, R.A. Compact and Highly Sensitive Bended Microwave Liquid Sensor Based on a Metamaterial Complementary Split-Ring Resonator. Appl. Sci. 2022, 12, 2144. https://doi.org/10.3390/app12042144

AMA Style

Mosbah S, Zebiri C, Sayad D, Elfergani I, Bouknia ML, Mekki S, Zegadi R, Palandoken M, Rodriguez J, Abd-Alhameed RA. Compact and Highly Sensitive Bended Microwave Liquid Sensor Based on a Metamaterial Complementary Split-Ring Resonator. Applied Sciences. 2022; 12(4):2144. https://doi.org/10.3390/app12042144

Chicago/Turabian Style

Mosbah, Said, Chemseddine Zebiri, Djamel Sayad, Issa Elfergani, Mohamed Lamine Bouknia, Samira Mekki, Rami Zegadi, Merih Palandoken, Jonathan Rodriguez, and Raed A. Abd-Alhameed. 2022. "Compact and Highly Sensitive Bended Microwave Liquid Sensor Based on a Metamaterial Complementary Split-Ring Resonator" Applied Sciences 12, no. 4: 2144. https://doi.org/10.3390/app12042144

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