Next Article in Journal
The Role of Ammonium Chloride in the Powder Thermal Diffusion Alloying Process on a Magnesium Alloy
Next Article in Special Issue
Material Application of a Transformer Box: A Study on the Electromagnetic Shielding Characteristics of Al–Ta Coating Film with Plasma-Spray Process
Previous Article in Journal
Surface Modification of Biomedical Titanium Alloy: Micromorphology, Microstructure Evolution and Biomedical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

TiO2 Nano Flowers Based EGFET Sensor for pH Sensing

1
Green Energy Technology Research Center, Kun Shan University, Tainan 710, Taiwan
2
Department of Electrical Engineering, Institute of Microelectronics, National Cheng Kung University, Tainan 701, Taiwan
3
Department of Electrical Engineering, Kun Shan University, Tainan 710, Taiwan
*
Authors to whom correspondence should be addressed.
Coatings 2019, 9(4), 251; https://doi.org/10.3390/coatings9040251
Submission received: 19 March 2019 / Revised: 11 April 2019 / Accepted: 11 April 2019 / Published: 15 April 2019
(This article belongs to the Special Issue Advances in Thin Film Transistors: Properties and Applications)

Abstract

:
In this study, pH sensors were successfully fabricated on a fluorine-doped tin oxide substrate and grown via hydrothermal methods for 8 h for pH sensing characteristics. The morphology was obtained by high-resolution scanning electron microscopy and showed randomly oriented flower-like nanostructures. The TiO2 nanoflower pH sensors were measured over a pH range of 2–12. Results showed a high sensitivity of the TiO2 nano-flowers pH sensor, 2.7 (μA)1/2/pH, and a linear relationship between IDS and pH (regression of 0.9991). The relationship between voltage reference and pH displayed a sensitivity of a 46 mV/pH and a linear regression of 0.9989. The experimental result indicated that a flower-like TiO2 nanostructure extended gate field effect transistor (EGFET) pH sensor effectively detected the pH value.

1. Introduction

The pH value, which is defined as the negative logarithm of hydrogen ion concentration, is a standardized scale for determining whether a solution is alkaline or acidic. The more acidic is the liquid, the greater the amount of hydrogen ions. Acid–base sensing has been used in numerous applications, such as water quality testing, food safety, agriculture, chemistry, and medical care, indicating its importance [1,2,3].
To date, pH sensors that are the most commonly used are pH test strips and meters. A pH test strip is a simple, low cost, and fast method, but its disadvantage is that the specific pH value cannot be accurately determined. On the contrary, a pH meter commonly uses glass electrodes, which are costly and fragile, proving difficult to maintain. On this basis, two types of typical sensing components are available. One is the ion-sensitive field-effect transistor (ISFET), and the other is the extended gate field effect transistor (EGFET) [4,5], which are used to measure the hydrogen ion concentration in the medium. ISFET is similar to the metal-oxide-semiconductor field-effect transistor (MOSFET). The difference is that the metal gate of the MOSFET is changed to an insulating layer (e.g., SiO2) as a sensing layer. To determine the pH value, the concentration of hydrogen ions, which change the surface potential and the transistor’s current, is identified. Different from ISFET, EGFET has an independent layer structure fabricated on the end of the signal line extended from the field-effect transistor gate electrode. The independent layer mostly uses metal oxide as a sensing membrane, which can attract the hydrogen ions in the solution. In comparison with ISFET, EGFET has several advantages: (1) the sensing region is separated from the gate-controlled region, which simply replaces the sensing film, so it can be developed as a disposable component; (2) it is easy to package due to its relatively simple structure; (3) the sensing film can be easily switched; and (4) there is no light interference [5].
Metal oxides, such as ZnO, CuO, SnO2, and WO3 have been widely used as sensing films in numerous studies [6,7,8,9]. TiO2 is a semiconductor, which has received considerable attention due to its superior chemical stability, including nontoxicity, acid and alkaline resistance, low cost, and is easy to process [10]. Accordingly, TiO2 has been utilized in gas sensors, UV photodetectors, memory resistors, and water treatment cleaning applications [11,12,13,14,15,16,17]. By contrast, the thin film-based sensing membrane has poor sensitivity. On the basis of the poor sensitivity, several studies have reported a nanostructure, which can enhance the component performance due to the large surface area [18]. The nanostructure can be fabricated by self-assembly growth processes, such as the hydrothermal method and sputtered and metal organic chemical vapor deposition (CVD). However, sputter and CVD methods should be used under a high-vacuum ambient environment. In addition, their source materials are expensive. Thus, the hydrothermal method is mostly used because of its advantages, such as low-cost fabrication, low-temperature process, and easy processing [19].
In this work, to increase the device sensitivity, we prepared a TiO2 nanostructure on a fluorine-doped tin oxide (FTO) substrate via the hydrothermal method. The TiO2 EGFET pH sensor was measured with a solution of pH 2–12. The device performance and the material analysis are also discussed herein.

2. Experiment

For the device preparation, 2 × 2 cm2 FTO glass was cleaned by acetone, isopropanol, and deionized (DI) water using an ultrasonicator (DELTA D150, Taipei, Taiwan). The FTO substrate that can be prepared in a strong acid environment also has small lattice mismatch with TiO2 [20]. Before cleaning, the 0.5 cm fluoroplastic adhesive tape sticks (Nitto Denko No.903UL, Osaka, Japan) were placed on the substrate and used as an electrode. Figure 1 shows the schematic diagram of the TiO2 EGFET pH sensor. The precursor solution was premixed with DI water (30 mL) and 36% hydrochloric acid (30 mL), added with titanium(IV) n-butoxide (>99%, 2 mL), and ultrasonicated for 20 min. The formation reaction of TiO2 is shown in the following equation [21]:
TiCl4 + H2O → HCl + Ti(IV) complex
2 Ti ( IV )   complex   dehydration   TiO 2
After completing the preparation, the precursor was placed in a Teflon autoclave, and the sample was positioned facing upward. Then, the Teflon autoclave was placed in a hot oven and heated at 150 °C for 8 h. Finally, the sample underwent argon flow annealing for 60 min at 300 °C. A standard buffer solution (Great & Best Co., Ltd., Taipei, Taiwan) was used in the sensing measurement at a range of pH 2–12. Characterization of the TiO2 nanostructure included X-ray diffraction (XRD, D8 Discover, Bruker, Billerica, MA, USA), high-resolution scanning electron microscopy (HR-SEM, HITACHI SU8000, Tokyo, Japan), and energy-dispersive spectroscopy (EDS), as discussed further below. Semiconductor parameter analyzer 4145B (Keysight Technologies, Santa Rosa, CA, USA) was set up to measure the current–voltage (IV) characteristic curves. The pH sensing electrode was connected to the gate (pin 6) of a commercial CD4007UB n-type MOSFET, and pins 7 and 8 corresponded to the source and drain, respectively. A commercial Ag/AgCl electrode was used as reference electrode. The distance between the reference and sensing electrodes was fixed at approximately 2 cm, and the process was performed in a dark room at room temperature. Figure 2 shows the schematic diagram of the measuring device.

3. Results and Discussion

Figure 3 shows the XRD pattern of synthesized TiO2 nanoflowers and the peaks at 2θ = 27.46°, 36.11°, 41.26°, 44.08°, 54.37°, 56.69°, 62.79°, 64.09°, 69.07°, and 69.85°, which correspond to (110), (101), (111), (210), (211), (220), (002), (310), (301), and (112) planes, respectively, indicating that the sample of TiO2 nanoflower was tetragonal rutile TiO2, which can be indexed to JCPDS card no. 21-1276. Figure 4 shows the top view of the HR-SEM image of the hydrothermal synthesis of TiO2 samples, revealing a crystalline flower-like and randomly oriented structure. Meanwhile, the EDS analyses detected only Ti and oxygen. The atomic percentages of Ti and oxygen were 43.28 at.% and 56.72 at.%, respectively, and no other elements were detected (insert in Figure 3).
Figure 5a shows the Vref-IDS transfer characteristic of the device, setting the MOSFET at a gate voltage of 0–3 V and at VDS = 0.3 V in pH 2–12 buffer solution. Vref decreased with an increase in the pH value due to the lower H+ ion concentrations in the buffer solution, decreasing the surface potential of the sensing electrode. VT(EGFET) can be expressed as follows [22]:
V T ( EGFET ) = V T ( MOSFET ) M q + E ref + X Sol φ
where VT(MOSFET) is the threshold voltage of MOSFET, M denotes the work function of the reference electrode, q represents the electron charge, E ref indicates the potential of the reference electrode, X Sol refers to the surface dipole potential of the buffer solution, and φ signifies the surface potential at the electrolyte/sensing film interface. Figure 5b shows the Vref value as a function of pH value. The pH sensitivity from the liner pH-Vref can be expressed as follows:
sensitivity = Δ V T Δ pH
The Vref values were 1.53, 1.66, 1.74, 1.84, 1.94, and 2 V for different pH values ranging from 2 to 12. The device sensitivity was 46 mV/pH and the high linear regression was 0.9989. Figure 6 shows the IDS-VDS output characteristic. The output characteristic of the device can be expressed as follows [23]:
I DS = W μ n C ox 2 L [ ( V ref V T ) 2 ]
where W and L are the channel width and length, respectively, C o x repesents the oxide capacitance per unit area, μ n denotes the electron mobility in the channel, and V T indicates the threshold voltage. IDS increased with the pH value from 12 to 2, indicating that the device can be modulated by various pH values. Figure 6b shows the IDS value as a function of pH value. The pH current sensitivity from the linear pH-IDS can be expressed as follows:
sensitivity = Δ I DS Δ pH
The IDS values were 1.25, 1.19, 1.14, 1.08, 1.03, and 0.99 for different pH values ranging from 2 to 12. The current sensitivity of the device was 2.7 (μA)1/2/pH, and the high linear regression was 0.9982. The sensing performance can be explained based on the site-binding model [24]. The surface potential depends on the pH of a buffer solution; thus, the sensing Ti-membrane has three different forms: negative (Ti-O), positive (Ti-OH2+), and neutral (Ti OH). Furthermore, a lower pH value indicates a greater concentration of H+ in the solution. Consequently, a high amount of H+ is adsorbed onto the sensing membrane. Table 1 displays the device sensitivity compared with other studies given the great specific surface area, resistance to high temperature, acid, and alkali, and high electrical conductivity of TiO2 nanoflowers; as such, the device exhibited more potential than other EGFET pH sensors. Dar et al. [25] reported that glucose is easily oxidized by dynamic oxygen species, and metal oxide layers facilitate the release of trapped electrons back to the conduction band of metal oxide. Oxygen is absorbed at the interface of metal oxide, ionized, and converted into dynamic oxygen species. Chen et al. [26] reported that uric acid (UA) can be easily oxidized in common electrodes in aqueous solutions, thereby modifying the detection potential for different concentrations. Hence, these sensors have promising prospects for detection.

4. Conclusions

In this study, pH sensors were successfully fabricated on FTO substrate and grown via hydrothermal methods at 8 h for pH sensing characteristics. Results indicated that a pH sensor with a TiO2 nanoflower can measure in the pH range of 2–12 buffer solution and the nanoflower growth has the best sensitivity and linear regression relationship between IDS and pH, at 2.7 (μA)1/2/pH and 0.9991, respectively. Moreover, the relationship among voltage reference, pH, and sensitivity was 46 mV/pH, and the linear regression was 0.9989. Thus, this new approach is not only expected to be useful for pH sensors but also to detect glucose or uric acid via an advanced multifunctional biosensor.

Author Contributions

Conceptualization, C.-C.Y. and K.-Y.C.; Methodology, C.-C.Y. and K.-Y.C.; Validation, C.-C.Y., K.-Y.C. and Y.-K.S.; Formal Analysis, C.-C.Y. and K.-Y.C.; Resources, C.-C.Y. and Y.-K.S.; Data Curation, C.-C.Y. and K.-Y.C.; Writing—Original Draft Preparation, C.-C.Y. and K.-Y.C.; Writing—Review and Editing, C.-C.Y., K.-Y.C. and Y.-K.S.

Funding

This research was supported by the Ministry of Science and Technology of Taiwan, ROC, under Contract Nos. MOST 104-2221-E-168-011-MY3 and 107-2221-E-006-185-MY3 and the Green Energy Technology Research Center, Department of Electrical Engineering, Kun Shan University, Tainan, Taiwan, from The Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education (MOE) in Taiwan.

Conflicts of Interest

The authors declare no conflicts of interests.

References

  1. Wang, Y.; Huang, T.; Liu, J.; Lin, Z.; Li, S.; Wang, R.; Ge, Y. Soil pH value, organic matter and macronutrients contents prediction using optical diffuse reflectance spectroscopy. Comput. Electron. Agric. 2015, 111, 69–77. [Google Scholar] [CrossRef]
  2. Honma, T.; Ohba, H.; Kaneko-Kadokura, A.; Makino, T.; Nakamura, K.; Katou, H. Optimal soil Eh, pH, and water management for simultaneously minimizing arsenic and cadmium concentrations in rice grains. Environ. Sci. Technol. 2016, 50, 4178–4185. [Google Scholar] [CrossRef] [PubMed]
  3. Schumacher, S.P.; Schurgers, L.J.; Vervloet, M.G.; Neradova, A. Influence of pH and phosphate concentration on the phosphate binding capacity of five contemporary binders. An in vitro study. Nephrology 2019, 24, 221–226. [Google Scholar] [CrossRef] [PubMed]
  4. Moser, N.; Lande, T.S.; Toumazou, C.; Georgiou, P. ISFETs in CMOS and emergent trends in instrumentation: A review. IEEE Sens. J. 2016, 16, 6496–6514. [Google Scholar] [CrossRef]
  5. Pullano, S.; Critello, C.; Mahbub, I.; Tasneem, N.; Shamsir, S.; Islam, S.; Greco, M.; Fiorillo, A.S. EGFET-based sensors for bioanalytical applications: A review. Sensors 2018, 18, 4042. [Google Scholar] [CrossRef]
  6. Qi, J.; Zhang, H.; Ji, Z.; Xu, M.; Zhang, Y. ZnO nano-array-based EGFET biosensor for glucose detection. Appl. Phys. A 2015, 119, 807–811. [Google Scholar] [CrossRef]
  7. Chen, P.Y.; Yin, L.T.; Cho, T.H. Optical and impedance characteristics of EGFET based on SnO2/ITO sensing gate. Life Sci. J. 2014, 11, 871–875. [Google Scholar]
  8. Guidelli, E.J.; Guerra, E.M.; Mulato, M. V2O5/WO3 mixed oxide films as pH-EGFET sensor: Sequential re-usage and fabrication volume analysis. ECS J. Solid State Sci. Technol. 2012, 1, N39–N44. [Google Scholar] [CrossRef]
  9. Chang, S.P.; Yang, T.H. Sensing performance of EGFET pH sensors with CuO nanowires fabricated on glass substrate. Int. J. Electrochem. Sci. 2012, 7, 5020–5027. [Google Scholar]
  10. Galiano, F.; Song, X.; Marino, T.; Boerrigter, M.; Saoncella, O.; Simone, S.; Faccini, M.; Chaumette, C.; Drioli, E.; Figoli, A. Novel photocatalytic PVDF/nano-TiO2 hollow fibers for environmental remediation. Polymers 2018, 10, 1134. [Google Scholar] [CrossRef]
  11. Galstyan, V. Porous TiO2-based gas sensors for cyber chemical systems to provide security and medical diagnosis. Sensors 2017, 17, 2947. [Google Scholar] [CrossRef]
  12. Liu, H.Y.; Sun, W.C.; Wei, S.Y.; Yu, S.M. Characterization of TiO2-based MISIM ultraviolet photodetectors by ultrasonic spray pyrolysis. IEEE Photonics Technol. Lett. 2016, 28, 637–640. [Google Scholar] [CrossRef]
  13. Chen, Y.; Li, L.; Yin, X.; Yerramilli, A.; Shen, Y.; Song, Y.; Bian, W.; Li, N.; Zhao, Z.; Qu, W.; et al. Resistive switching characteristics of flexible TiO2 thin film fabricated by deep ultraviolet photochemical solution method. IEEE Electron Device Lett. 2017, 38, 1528–1531. [Google Scholar] [CrossRef]
  14. Lazar, M.; Varghese, S.; Nair, S. Photocatalytic water treatment by titanium dioxide: Recent updates. Catalysts 2012, 2, 572–601. [Google Scholar] [CrossRef]
  15. Joshi, N.; Hayasaka, T.; Liu, Y.; Liu, H.; Oliveira, O.N., Jr.; Lin, L. A review on chemiresistive room temperature gas sensors based on metal oxide nanostructures, graphene and 2D transition metal dichalcogenides. Mikrochim. Acta 2018, 185, 213. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, X.; Ma, T.; Pinna, N.; Zhang, J. Two-dimensional nanostructured materials for gas sensing. Adv. Funct. Mater. 2017, 27, 1702168. [Google Scholar] [CrossRef]
  17. Zhang, J.; Liu, X.; Neri, G.; Pinna, N. Nanostructured materials for room-temperature gas sensors. Adv. Mater. 2016, 28, 795–831. [Google Scholar] [CrossRef] [PubMed]
  18. Zhang, Q.; Liu, W.; Sun, C.; Zhang, H.; Pang, W.; Zhang, D.; Duan, X. On-chip surface modified nanostructured ZnO as functional pH sensors. Nanotechnology 2015, 26, 355202. [Google Scholar] [CrossRef] [PubMed]
  19. Zhang, J.; Li, J.; Chen, P.; Rehman, F.; Jiang, Y.; Cao, M.; Zhao, Y.; Jin, H. Hydrothermal growth of VO2 nanoplate thermochromic films on glass with high visible transmittance. Sci. Rep. 2016, 6, 27898. [Google Scholar] [CrossRef]
  20. Issar, S.; Poddar, P.; Mehra, N.C.; Mahapatro, A.K. Growth of flower-like patterns of TiO2 nanorods over FTO substrate. Integr. Ferroelectr. 2017, 184, 166–171. [Google Scholar] [CrossRef]
  21. Kumar, A.; Madaria, A.R.; Zhou, C. Growth of aligned single-crystalline rutile TiO2 nanowires on arbitrary substrates and their application in dye-sensitized solar cells. J. Phys. Chem. C 2010, 114, 7787–7792. [Google Scholar] [CrossRef]
  22. Das, A.; Ko, D.H.; Chen, C.H.; Chang, L.B.; Lai, C.S.; Chu, F.C.; Chow, L.; Lin, R.M. Highly sensitive palladium oxide thin film extended gate FETs as pH sensor. Sens. Actuators B 2014, 205, 199–205. [Google Scholar] [CrossRef]
  23. Bergveld, P. The impact of MOSFET-based sensors. Sens. Actuators 1985, 8, 109–127. [Google Scholar] [CrossRef] [Green Version]
  24. Yusof, K.A.; Abdul Rahman, R.; Zulkefle, M.A.; Herman, S.H.; Abdullah, W.F.H. EGFET pH sensor performance dependence on sputtered TiO2 sensing membrane deposition temperature. J. Sens. 2016, 2016, 7594531. [Google Scholar] [CrossRef]
  25. Dar, G.N.; Umar, A.; Zaidi, S.A.; Baskoutas, S.; Kim, S.H.; Abaker, M.; Al-Hajry, A.; Al-Sayari, S.A. Fabrication of highly sensitive non-enzymatic glucose biosensor based on ZnO nanorods. Sci. Adv. Mater. 2011, 3, 901–906. [Google Scholar] [CrossRef]
  26. Chen, X.; Wu, G.; Cai, Z.; Oyama, M.; Chen, X. Advances in enzyme-free electrochemical sensors for hydrogen peroxide, glucose, and uric acid. Microchim. Acta 2013, 181, 689–705. [Google Scholar] [CrossRef]
  27. Ali, G.M. Interdigitated extended gate field effect transistor without reference electrode. J. Electron. Mater. 2017, 46, 713–717. [Google Scholar] [CrossRef]
  28. Guerra, E.M.; Mulato, M. Synthesis and characterization of vanadium oxide/hexadecylamine membrane and its application as pH-EGFET sensor. J. Sol-Gel Sci. Technol. 2009, 52, 315–320. [Google Scholar] [CrossRef]
  29. Zulkefle, M.A.; Rahman, R.A.; Rusop, M.; Abdullah, W.F.H.; Herman, S.H. Porous TiO2 thin film for EGFET pH sensing application. Int. J. Eng. Technol. 2018, 7, 112–114. [Google Scholar]
  30. Li, J.Y.; Chang, S.P.; Chang, S.J.; Tsai, T.Y. Sensitivity of EGFET pH sensors with TiO2 nanowires. ECS Solid State Lett. 2014, 3, P123–P126. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the TiO2 extended gate field effect transistor (EGFET) sensor.
Figure 1. Schematic diagram of the TiO2 extended gate field effect transistor (EGFET) sensor.
Coatings 09 00251 g001
Figure 2. EGFET measurement system.
Figure 2. EGFET measurement system.
Coatings 09 00251 g002
Figure 3. XRD pattern of synthesized TiO2 nano-flowers.
Figure 3. XRD pattern of synthesized TiO2 nano-flowers.
Coatings 09 00251 g003
Figure 4. (a) Top view of HR-SEM image of synthesized TiO2 nano-flowers, which show the randomly oriented flower-like nanostructures. (b) The high-magnified HR-SEM image of TiO2 nano-flowers.
Figure 4. (a) Top view of HR-SEM image of synthesized TiO2 nano-flowers, which show the randomly oriented flower-like nanostructures. (b) The high-magnified HR-SEM image of TiO2 nano-flowers.
Coatings 09 00251 g004
Figure 5. (a) IDS-versus-Vref curve under various pH buffer solutions. (b) Linearity of the Vref of TiO2 EGFET sensor.
Figure 5. (a) IDS-versus-Vref curve under various pH buffer solutions. (b) Linearity of the Vref of TiO2 EGFET sensor.
Coatings 09 00251 g005
Figure 6. (a) IDS-versus-VDS curve under various pH buffer solutions. (b) Linearity of the IDS of TiO2 EGFET sensor.
Figure 6. (a) IDS-versus-VDS curve under various pH buffer solutions. (b) Linearity of the IDS of TiO2 EGFET sensor.
Coatings 09 00251 g006
Table 1. Comparison of other metal oxide-based extended gate field effect transistor (EGFET) pH sensors.
Table 1. Comparison of other metal oxide-based extended gate field effect transistor (EGFET) pH sensors.
SampleFabrication MethodSensitivity (mV/pH)Ref.
ZnO thin filmsol–gel26.5[27]
V2O5/hadHydrothermal38.1[28]
Porous TiO2Hydrothermal19.3[29]
TiO2 nanowiresE-beam evaporator32.65[30]
TiO2 nanoflowerHydrothermal46This work

Share and Cite

MDPI and ACS Style

Yang, C.-C.; Chen, K.-Y.; Su, Y.-K. TiO2 Nano Flowers Based EGFET Sensor for pH Sensing. Coatings 2019, 9, 251. https://doi.org/10.3390/coatings9040251

AMA Style

Yang C-C, Chen K-Y, Su Y-K. TiO2 Nano Flowers Based EGFET Sensor for pH Sensing. Coatings. 2019; 9(4):251. https://doi.org/10.3390/coatings9040251

Chicago/Turabian Style

Yang, Chih-Chiang, Kuan-Yu Chen, and Yan-Kuin Su. 2019. "TiO2 Nano Flowers Based EGFET Sensor for pH Sensing" Coatings 9, no. 4: 251. https://doi.org/10.3390/coatings9040251

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