i Synthesis of zirconium disulphide nanomaterials and their nanocomposites with radially aligned nanorutile and polyaniline for room temperature sensing of volatile organic compounds By Paul Olawale Fadojutimi Student no: 2165101 A thesis submitted to the Faculty of Science, University of the Witwatersrand, Johannesburg, in partial fulfilment of the requirements for the award of the degree of PhD in Chemistry. Supervisor: Dr John Moma Co-Supervisor: Dr Siziwe Gqoba Co-Supervisor: Dr Zikhona N. Tetana University of the Witswatersrand, Johannesburg, August 2022 ii Declaration I declare that this thesis is my own unaided work. It is being submitted for the degree of Doctor of Philosophy in Science at the University of the Witwatersrand, Johannesburg. It has not been submitted before for any degree or examination at any other university. (Signature of candidate) On this 17th day of August 2022 iii Abstract Integration of 2D nanomaterials with a polymer or semiconductor metal oxide could help in the development of low-cost sensors for rapid detection of volatile organic compounds (VOCs) at room temperature. This study focuses on the fabrication of robust room temperature sensors of pristine radially aligned nanorutile, and zirconium disulphide/polyaniline (PANI) nanocomposites for chemical sensing of VOCs. ZrS2 was fabricated using both bottom-up and top-down methods of synthesis. Heat up and hot injection methods were employed to fabricate arrays of morphologies of ZrS2 nanomaterials using the colloidal method. However, the nanomaterials showed high oxophilicity which was confirmed by both XRD and XPS. The XPS peak of S2p was conspicuously absent while the peak Zr3d was very noticeable in the XPS spectra. XPS and EDS measurements indicated replacement of sulphur atom by the O atom on the surface of the nanomaterials. The stability study showed the nanomaterials were not stable in ambient environment. Nanoparticles of 11 nm and few layered nanosheets were obtained when bulk crystal samples of ZrS2 were exfoliated in cyclohexyl-2-pyrrolidone and N-methyl (-2-) pyrrolidone. Isopropanol served as a green solvent for the exfoliation of few-layered ZrS2 from the bulk crystal sample compared to amide solvents which are not environmentally friendly. However, the pristine ZrS2 nanomaterials could not sense VOCs at room temperature, this could be as a result of low conductivity and number of layers of the nanomaterials obtained using nanomaterials exfoliated in isopropanol. The sensitivity of raw PANI was greatly enhanced with loading of ZrS2 nanomaterials. The sensor displayed responses of 0.43, 0.58, 1.04 and 0.34% which correspond to methanol, ethanol, isopropanol, and acetone vapours respectively. The relatively better responses of the sensor were credited to the synergistic effect of ZrS2/PANI composite structure. The sensor showed good response to low concentrations (7.7 ppm, 11 ppm, 5.8 ppm and 6.1 ppm) which correspond to methanol, ethanol, isopropanol, and acetone respectively. The sensor was more sensitive to isopropanol compared to other alcohols tested for in this work. The behaviour of the sensor changed from p-type to n-type on exposure to ethanol vapour at elevated relative humidity. The sensor displayed good sensitivity, reproducibility, rapid response and recovery times towards alcohols and stability over 60 days. iv The hierarchical morphology, high surface area, high porosity and humidity contributed immensely to the titania sensor in the sensing of VOCs at room temperature. The TiO2 sensor showed high sensitivity with responses of -38.27, -86.75, -9.83 and 1.24% which correspond to methanol, ethanol, isopropanol, and acetone respectively. The sensor is more sensitive to ethanol gas compared to other chemical vapours tested. The sensor displayed good sensitivity, reproducibility, rapid response and recovery times towards alcohols and stability over 45 days. The surface area of the nanorutile decreased by 35% on loading of ZrS2, this could the main reason there was no response observed when the nanocomposite of TiO2-ZrS2 was tested for chemical sensing at room temperature. The active sites for adsorption of the vapour were not available probably due to the covering of pores of the nanorutile as well as low conductivity of ZrS2 at room temperature. Both sensors of nanocomposite of zirconium disulphide and polyaniline as well as titania could find application in breath analysers since the least detection for a breath analyser is reported to be 200 ppm. v Dedication This work is dedicated to my late parents Mr and Mrs Ebenezer Olakunle Fadojutimi who valued good education and ensured they gave me the best legacy a parent can give to a child. vi Acknowledgements I would like to convey my sincere appreciation to everyone who made this work a success: ❖ Unto God be the glory, great things He has done. I would like to thank God Almighty, the Elshadai for provisions and giving me wisdom, good health, and vigour to complete this research. ❖ I am indeed grateful to all my supervisors, Dr John Moma, Dr Siziwe Gqoba, Dr Zikhona Tetana and Prof Nosipho Moloto for their guidance, advice and support in this research. Thank you so much for your constructive criticism and for your painstaking effort going through my thesis. I am indeed privileged to learn from the best. ❖ A big thanks to all my colleagues in CATMAT group and most especially everyone in labs 117 and 330 for their support and friendly environment provided so far. I would like to thank the following colleagues especially: Dr Rudo Sithole, Dr Pumza Mente, Dr Tumelo Phaahlamohlaka, Dr Victor Mashindi, Mr Clinton Masemola, Miss Boipelo Mathe and Mr Themba Ntuli. ❖ Special thanks to every staff member of the Microscopy and Microanalysis Unit (MMU) for the XRD, SEM and TEM training. My special appreciation goes to Dr Rudolph Erasmus for his assistance in Raman spectroscopy analyses and Dr Sanele Nyembe from MINTEK who assisted with AFM analyses. ❖ Many thanks to Wits University for funding this research work. ❖ Lastly, kudos to my lovely wife Oluwayemisi and my wonderful children Oluwakoretimi and Irenitemi Fadojutimi for your support despite my absence from home. vii Presentations 1. Colloidal synthesis of zirconium disulphide (oral presentation), Catalysis and Materials Science group (March 2021). 2. Nanocomposite of rutile titania and zirconium disulphide for room temperature sensing of acetone vapour (Poster presentation) National Young Chemist’s symposium (July 2021). 3. Synthesis of zirconium disulphide nanostructures for room temperature sensing of acetone vapours (poster presentation), Wits Nanoscience Young researcher’s symposium (October 2021). viii Publications 1. Paul Fadojutimi, Siziwe Gqoba, Zikhona Tetana and John Moma. Transition metal dichalcogenides in photocatalytic water splitting. (2022) Journal of Catalysts. Manuscript published. 2. Siziwe S. Gqoba, Rafael Rodrigues, Sharon Lerato Mphahlele , Zakhele Ndala , Mildred Airo , Paul Olawale Fadojutimi , Ivo A. Hümmelgen , Ella C. Linganiso , Makwena J. Moloto and Nosipho Moloto. Hierarchical Nanoflowers of Colloidal WS2 and Their Potential Gas Sensing Properties for Room Temperature. (2021). J. of Processes, 9, 1491. 3. Paul Fadojutimi, Zikhona Tetana, John Moma, Nosipho Moloto and Siziwe Gqoba Colloidal synthesis of zirconium disulphide nanostructures and its stability against oxidation. (2022). Journal of Chemistry Select. Manuscript published. Manuscripts to be submitted. 1. Paul Fadojutimi, Siziwe Gqoba, Zikhona Tetana, and John Moma. Hierarchical dandelion rutile titania and their potential gas sensing properties for room temperature sensing methanol and ethanol. 2. Paul Fadojutimi, Siziwe Gqoba, Zikhona Tetana, John Moma and Nosipho Moloto. Nanocomposite of polyaniline and zirconium disulphide for room temperature sensing of primary alcohol. 3. Paul Fadojutimi, Siziwe Gqoba, Zikhona Tetana, Nosipho Moloto, and John Moma. Nanocomposite of polyanile and titania for room temperature sensing of primary alcohols. 4. Paul Fadojutimi, Siziwe Gqoba, Zikhona Tetana, John Moma and Nosipho Moloto. Few layered synthesis and characterization of zirconium disulphide for optoelectronic application. ix Table of Contents Declaration…………………………………………………………………………………….ii Abstract……………………………………………………………………………………….iii Dedication…………………………………………………………………………………….v Acknowledgements…………………………………………………………………………..vi Presentations…………………………………………………………………………………vii Publication…………………………………………………………………………………..viii Table of contents…………………………………………………………………………….ix List of figures………………………………………………………………………………..xv List of tables………………………………………………………………………………………...xxi Abbreviation Description………………………………………………………………………………….xxii Chapter 1: Introduction and motivation……………………………………………………...1 1.1 Background and motivation …………………………………………………………….1 1.2 Problem statement……………………………………………………………………….3 1.3 Aim and objectives………………………………………………………………………4 1.4 Thesis outline……………………………………………………………………………4 1.5 References……………………………………………………………………………….6 Chapter 2: Literature review………………………………………………………………….8 2.1 Transition metal dichalcogenides for chemical sensing………………………………….8 2.2. Exfoliation……………………………………………………………………………….9 2.3 Gas-Phase method of metal powder production…………………………………………12 2.3.1 Chemical vapour deposition……………………………………………………………12 2.3.2 Chemical vapour transport …………………………………………………………….13 x 2.4 Wet chemical synthesis…………………………………………………………………...14 2.4.1 Colloidal synthesis …………………………………………………………………….14 2.4.2 Hydrothermal or solvothermal synthesis……………………………………………….19 2.4.3 Sol-gel method………………………………………………………………………….20 2.5 Gas sensing mechanism in TMDCs………………………………………………………20 2.6 Attributes of a gas sensor…………………………………………………………………21 2.6.1 Response ……………………………………………………………………………….21 2.6.2 Response and Recovery Time …………….. .. …………………………………………22 2.6.3 Selectivity/Specificity ………………………………………………………………...22 2.6.4 Detection limits…………………………………………………………………………22 2.6.5 Stability ………………………………………………………………………………...23 2.6.6 Sensitivity and grain size……………………………………………………………….23 2.7 Factors controlling sensitivity…………………………………………………………….24 2.7.1 Receptor………………………………………………………………………………...24 2.7.2 Tranducer……………………………………………………………………………….24 2.7.3 Utility factor…………………………………………………………………………….25 2.8 Gas sensing properties of 2D TMDCs and other 2D materials…………………………..26 2.8.1 Metal doping……………………………………………………………………………30 2.8.2 Metal Oxide for sensing VOCs …………………………………………………………30 2.8.3 TMDCs with metal oxide composite …………………………………………………..32 2.8.4 PANI for chemical sensing……………………………………………………………..34 2.9 References………………………………………………………………………………..35 xi CHAPTER 3: colloidal synthesis of zirconium disulphide and its stability against oxidation……………………………………………………………………………………...47 3.1 Introduction………………………………………………………………………………48 3.2 Experimental……………………………………………………………………………..50 3.2 Chemicals and materials…………………………………………………………………50 3.2.2 Synthesis……………………………………………………………………………….50 3.2.3 Heat up Method………………………………………………………………………..50 3.2.4 Hot injection method…………………………………………………………………..50 3.3 Material Characterization………………………………………………………………..51 3.4 Results and Discussion…………………………………………………………………..51 3.5 Conclusion……………………………………………………………………………….63 3.6 References……………………………………………………………………………….64 CHAPTER 4: synthesis of radially aligned nanorutile and its application for chemical sensing of alcohols and acetone……………………………………………………………………...67 4.1 Introduction……………………………………………………………………………...67 4.2 Materials and methods…………………………………………………………………..69 4.2.1 Chemicals……………………………………………………………………………..69 4.2.2 Synthesis of radially aligned nanorutile (RANR)…………………………………….69 4.2.3 Sensor fabrication ……………………………………………………………………70 4.2.4 Gas sensing measurement …………………………………………………………....70 4.3 Characterization of RANR……………………………………………………………...71 4.4 Results…………………………………………………………………………………..72 4.4.1 PXRD OF RANR……………………………………………………………………..72 xii 4.4.2 SEM analysis………………………………………………………………………….73 4.4.3 TEM analysis………………………………………………………………………….74 4.4.4 UV-Vis Spectra analysis……………………………………………………………….76 4.4.5 Thermal analysis of RANR……………………………………………………………78 4.4.6 BET analysis…………………………………………………………………………...78 4.5 Gas sensing performance ………………………………………………………………..80 4.6 Methanol vapour sensing………………………………………………………………...81 4.7 Ethanol sensing…………………………………………………………………………..85 4.8 Isopropanol………………………………………………………………………………94 4.9 Acetone…………………………………………………………………………………...96 4.10 Conclusion……………………………………………………………………………. 100 4.11 References…………………………………………………………………………………..101 Chapter 5: Synthesis and characterization of few-layered zirconium disulphide via liquid exfoliation…………………………………………………………………………………...106 5.1 Introduction ……………………………………………………………………………..106 5.2 Chemicals and materials………………………………………………………………...107 5.3 Exfoliation of bulk ZrS2…………………………………………………………………107 5.4 Characterization…………………………………………………………………………108 5.5 Results and Discussion………………………………………………………………….109 5.5.1 XRD analysis……………………………………………………………………….....109 5.5.2 UV-Vis spectra analysis…………………………………………………………......110 5.5.3 Raman analysis………………………………………………………………………..113 xiii 5.5.4 AFM analysis…………………………………………………………………………114 5.5.5 SEM analysis………………………………………………………………………….116 5.5.6 TEM analysis………………………………………………………………………….117 5.5.7 Energy dispersive x-ray spectroscopy (EDS) analysis………………………………..118 5.5.8 Thermogravimetry analysis…………………………………………………………...119 5.5.9. XPS analysis………………………………………………………………………….119 5.6 Conclusion……………………………………………………………………………....123 5.7 References………………………………………………………………………………124 Chapter 6: Polyaniline and zirconium disulphide nanocomposite and its application for chemical sensing of alcohols and acetone…………………………………………………..127 6.1 Introduction……………………………………………………………………………..127 6.2 Chemicals and materials………………………………………………………………...128 6.3 Methods………………………………………………………………………………....129 6.3.1 Synthesis of PANI…………………………………………………………………......129 6.3.2 Synthesis of a nanocomposite of zirconium disulphide and polyaniline………………130 6.3.3 Sensor fabrication……………………………………………………………………..130 6.3.4 Gas sensing measurement…………………………………………………………......130 6.4 Characterization…………………………………………………………………………131 6.5 Results and discussions………………………………………………………………….131 6.5.1 XRD analysis………………………………………………………………………….129 6.5.2 TEM analysis………………………………………………………………………….130 6.5.3 BET analysis…………………………………………………………………………..132 6.6 Gas sensing of PANI-ZrS2 nanocomposite………………………………………………136 xiv 6.6.1 Primary alcohols and acetone sensing properties……………………………………..135 6.7 Relative Humidity……………………………………………………………………….150 6.8 Conclusion………………………………………………………………………………151 6.9 References ………………………………………………………………………………152 Chapter 7: Titania-zirconium disulphide nanocomposite for sensing acetone vapour……..155 7.1 Introduction:…………………………………………………………………………….155 7.2 Chemicals and materials………………………………………………………………...157 7.3 Methods…………………………………………………………………………………157 7.3.1 Synthesis of nanocomposite of titania with zirconium disulphide ……………………157 7.3.2 Sensor fabrication……………………………………………………………………..157 7.3.3 Gas sensing measurement……………………………………………………………..157 7.4 Characterization…………………………………………………………………………158 7.5 Results and discussions…………………………………………………………………158 7.5.1 XRD analysis…………………………………………………………………………158 7.5.2 Raman analysis……………………………………………………………………….159 7.5.3 TEM analysis…………………………………………………………………………160 7.5.4 UV-Vis spectroscopy…………………………………………………………………161 7.5.5 Energy dispersive x-ray spectroscopy (EDS) analysis……………………………….162 7.5.6 BET analysis………………………………………………………………………….163 7.5.7 TGA analysis…………………………………………………………………………164 7.6 Gas sensing performance of TiO2-ZrS2………………………………………………...165 7.7 Conclusion………………………………………………………………………………167 7.8 References………………………………………………………………………………168 xv Chapter 8: General conclusions and recommendations for future studies………………….171 8.1 General conclusions…………………………………………………………………….171 8.2 Recommendations………………………………………………………………………172 Appendix A…………………………………………………………………………………174 List of Figures Figure 2.1: Schematic illustration of sensing mechanism of TMDCs……………………….21 Figure 2.2: Schematic illustration highlighting the importance of grain size……………….24 Figure 2.3: Three factors determining the response of a semiconductor gas sensor………...25 Figure 3.1: PXRD pattern of ZrS2 synthesized via the heat up and injection methods……..52 Figure 3.2: SEM micrographs of ZrS2 nanomaterials synthesized by the heat up method…54 Figure 3.3: SEM micrographs of ZrS2 nanomaterials synthesized by the hot injection method……………………………………………………………………………………….54 Figure 3.4: TEM images of ZrS2 synthesized by the heat up method……………………….55 Figure 3.5: TEM images of ZrS2 synthesized by the hot injection method………………....55 Figure 3 6: (a) UV-Vis spectrum and (b) band gap of as-synthesized ZrS2 by the heat up method……………………………………………………………………………………….56 Figure 3.7: (a) UV-Vis spectrum and (b) band gap of as-synthesized ZrS2 by the injection method……………………………………………………………………………………….57 Figure 3.8: FT1R spectra of (a) raw OA and (b) as-synthesized ZrS2 by the heat up method……………………………………………………………………………………….58 Figure 3.9: TGA and DTG thermograms of ZrS2 synthesized by the heat up method……...59 Figure 3.10: (a) and (b) EDS spectra of ZrS2 synthesized by the heat up method…………..60 Figure 3.11: High resolution spectra with focus on core levels C1s, O1s, N1s, valence, Zr3d and XPS survey spectrum of ZrS2 synthesized by heat up method…………………………61 xvi Figure 3.12: Time-dependent study of ZrS2 nanomaterials with XRD (a) ZrS2 synthesized by heat up method and (b) synthesized by hot injection method……………………………….62 Figure 4.1: Schematic diagram depicting the synthesis of RANR…………………………..70 Figure 4.2: Gas sensing set up……………………………………………………………….71 Figure 4.3: PXRD of radially aligned nanorutile……………………………………………73 Figure 4.4: (a) and (b) show the SEM micrographs of the synthesized RANR………… ….74 Figure 4.5: (a) and (b) TEM images of RANR ……………………………………………..75 Figure 4.6: Mechanism of the synthesis of RANR………………………………………….75 Figure 4.7: Particle size distribution of as-synthesized RANR……………………………..76 Figure 4.8: The solid-state UV-Vis DRS plot of RANR……………………………………77 Figure 4.9: The band-gap energy (hv) of RANR……………………………………………77 Figure 4.10: The thermogravimetric and derivative thermogravimetry analysis curves of RANR……………………………………………………………………………………….78 Figure 4.11: Nitrogen adsorption/desorption isotherms of RANR…………………………79 Figure 4.12: BJH pore size distributions of the RANR…………………………………….79 Figure 4.13: Response and recovery curves of RANR to methanol at different concentrations……………………………………………………………………………….82 Figure 4.14: Normalized response of the RANR sensor as a function of methanol vapour concentrations……………………………………………………………………………….83 Figure 4.15: Repeatability response curves for methanol…………………………………..84 Figure 4.16: Static response characteristic of RANR based sensor towards methanol...85 Figure 4.17: Static response and recovery curves of RANR to ethanol at different concentrations………………………………………………………………………………86 Figure 4.18: Static responses curves of RANR sensor measured under series of ethanol vapour concentration………………………………………………………………………..87 xvii Figure 4.19: Normalized response of the RANR sensor as a function of methanol vapour concentrations ………………………………………………………………………………87 Figure 4.20: Repeatability response curves for ethanol…………………………………….88 Figure 4.21: Transient response characteristic of RANR based sensor toward 385 ppm of ethanol vapour………………………………………………………………………………..89 Figure 4.22: Response of RANR towards 77 ppm of ethanol vapour under different humidity conditions at 10, 33,51, 55, 64, 76 and 96%............................................................................90 Figure 4.23: Stability of RANR sensor towards 77 ppm of ethanol vapour within 45 days…………………………………………………………………………………………...94 Figure 4.24: Static responses curves of RANR sensor measured under series of isopropanol vapour concentrations………………………………………………………………………...95 Figure 4.25: Normalized response of the RANR sensor as a function of isopropanol vapour concentrations………………………………………………………………………………...96 Figure 4.26: Static response and recovery curves of RANR toward acetone………………….97 Figure 4.27: Normalized response of the RANR sensor as a function of acetone vapour concentrations………………………………………………………………………………...98 Figure 4.28: Static responses curves of RANR sensor measured under series of acetone vapour concentrations………………………………………………………………………..............99 Figure 4.29: Repeatability response curves towards acetone vapour………………………..100 Figure 5.1: Exfoliation of bulk crystal of zirconium disulphide…………………………….108 Figure 5.2: XRD of exfoliated ZrS2 nanomaterials…………………………………………110 Figure 5.3: UV-Vis absorption of ZrS2……………………………………………………...111 Figure 5.4: UV-Vis absorption of ZrS2 on exposure to ambient environment for five days…112 Figure 5.5: Tauc plot of ZrS2 (a) Bulk (b) IPA exfol. (c) NMP exfol. (d) CHP exfol………113 Figure 5.6: Raman spectra of bulk and exfoliated nanosheets of ZrS2 with different thickness…………………………………………………………………………………….114 xviii Figure 5.7: AFM micrographs, height and spectrum of exfoliated ZrS2 nanosheets (a) ethanol (b) IPA (c) methanol and (d) water…………………………………………………………..115 Figure 5.8: SEM images of ZrS2 (a) and (b) bulk sample (c) IPA exfoliated (d) H2O exfoliated (e) MeOH exfoliated (f) EtOH exfoliated (g) CHP exfoliated (h) NMP exfoliated………..116 Figure 5.9: TEM and HRTEM images of ZrS2 (Ai) and (Aii) CHP exfoliated samples (Bi) and (Bii) NMP exfoliated samples (Ci) and (Cii) IPA exfoliated samples (di) and (Dii) H2O exfoliated samples (Ei) and (Eii) MeOH exfoliated samples F (i) and F (ii) EtOH exfoliated sample………………………………………………………………………………………117 Figure 5.10: EDS of ZrS2 (a) bulk sample (b) NMP sample (c) CHP sample (d) MeOH sample (e) EtOH sample (f) IPA sample (g) H2O sample…………………………………..118 Figure 5.11: Thermogravimetry and derivative thermogravimetry curves of bulk, CHP and IPA exfoliated sample……………………………………………………………………………119 Figure 5.12: High resolution core level spectra of NMP exfoliated ZrS2 nanosheets with focus on O1S, C1S, S2P+Zr3d, N1s,valence and survey spectrum……………………………….121 Figure 5.13: High resolution core level spectra of CHP exfoliated ZrS2 nanosheets with focus on O1S, C1S, S2P+Zr3d, N1s,valence and survey spectrum………………………………..122 Figure 5.14: High resolution core level spectra of IPA exfoliated ZrS2 nanosheets with focus on O1S, C1S, S2P+Zr3d, N1s,valence and survey spectrum……………………………….123 Figure 6.1: Schematic diagram of synthesis of PANI………………………………………..129 Figure 6.2: XRD pattern of PANI-ZrS2 nanocomposite…………………………………….132 Figure 6.3: TEM image of PANI…………………………………………………………….133 Figure 6.4: TEM images of PANI-ZrS2 nanocomposite……………………………………..134 Figure 6.5: (A) Nitrogen adsorption-desorption isotherm for PANI-ZrS2 (B): BJH pore size distributions of PANI-ZrS2………………………………………………………………….135 Figure 6.6: (A) Nitrogen adsorption-desorption isotherm for PANI- ZrS2 (B): BJH pore size distributions of PANI……………………………………………………………………….135 Figure 6.7: Schematic illustration of sensing mechanism of PANI-ZrS2…………………..137 xix Figure 6.8: Dynamic response of PANI-ZrS2 towards methanol vapour at RT…………….138 Figure 6.9: (A) Dynamic response curves of PANI-ZrS2 towards methanol vapour at RT (B) Response curves of the PANI-ZrS2 sensor taken under various methanol vapour concentration………………………………………………………………………………..139 Figure 6.10: Normalized response of the PANI-ZrS2 sensor towards methanol, ethanol, and isopropanol…………………………………………………………………………………140 Figure 6.11: Typical response and recovery curves of PANI-ZrS2 sensor to ethanol at 77 ppm and 385 ppm respectively………………………………………………………………….141 Figure 6.12: Dynamic response and recovery curves of PANI-ZrS2 towards ethanol vapour at RT…………………………………………………………………………………………..143 Figure 6.13 Response and recovery characteristic curves towards methanol ……………..144 Figure 6.14: Repeatability of the PANI-ZrS2 sensor to 385 ppm of ethanol vapour at RT…………………………………………………………………………………………..145 Figure 6.15: Dynamic response and recovery curves of PANI-ZrS2 towards isopropanol vapour at RT……………………………………………………………………………………….146 Figure 6.16: Comparison of PANI-ZrS2 sensor towards methanol, ethanol and isopropanol vapour at RT……………………………………………………………………………….147 Figure 6.17: Static response and recovery curves of PANI-ZrS2 towards acetone vapour at RT………………………………………………………………………………………… 148 Figure 6.18 :Normalized response of the PANI-ZrS2 sensor towards acetone vapour…….149 Figure 6.19 :Response of PANI-ZrS2 towards (A) 111 ppm of methanol vapour; (B) 77 ppm of ethanol vapour under different humidity conditions at 33, 46,58, 64 and 76% for methanol and at 33, 51, 58, 75 and 96% for ethanol……………………………………………………...151 Figure 7.1: PXRD of (a) nanocomposite of TiO2-ZrS2 and (b) TiO2 ……………………..159 Figure 7.2: Raman of RANR and nanocomposite of ZrS2 and with ZrS2-TiO2…………..160 Figure 7.3: TEM of TiO2-ZrS2 nanocomposite…………………………………………...161 xx Figure 7.4: HRTEM of TiO2-ZrS2 nanocomposite…………………………………………161 Figure 7.5: DRS-UV-Vis of (a) TiO2-ZrS2 and (b) TiO2……………………………….......162 Figure 7.6: EDS of TiO2-ZrS2 nanocomposite……………………………………………...163 Figure 7.7: (a) Nitrogen adsorption/desorption isotherms of TiO2-ZrS2 nanocomposite; and (b) BJH pore size distribution of the TiO2-ZrS2 nanocomposite……………………………164 Figure 7.8: TGA of pristine RANR and TiO2-ZrS2 nanocomposite………………………...165 Figure 7.9: Static response of pristine RANR towards 122 ppm acetone vapour at RT…………………………………………………………………………………………...166 Figure 7.10: Static response of TiO2-ZrS2 towards 610 ppm acetone vapour at RT……167 xxi List of Tables Table 1.1: Classification of volatile organic pollutants ……………………………………….1 Table 2.1: TMDCs, metal oxide and its nanocomposite for chemical sensing………………..28 Table 2.2: Comparison of 2D TMDCs with Metal oxide-based sensors………………………33 Table 4:1 Comparison of BET of hierarchical titania…………………………………………80 Table 4.2: Comparison of different MOXs towards VOCs sensing………………………….92 Table 6.1: Group IVB TMDCs and TMDCs nanocomposite with polymers for VOC sensing………………………………………………………………………………………149 Table 7.1: The BET comparison between pristine RANR and nanocomposite of TiO2- ZrS2…………………………………………………………………………………………163 xxii Abbreviation Description Ammonium persulphate APS Polyanilime PANI Transition metal dichalcogenides TMDCs Interdigitated electrodes IDEs Metal oxide semiconductors MOXs Room temperature RT N,N dimethylformamide DMF N-methyl polypyrrolidone NMF Part per million ppm Relative humidity RH Volatile organic compounds VOCs Quantum dots QDs Polyvinylpyrrolidone PVP Printed-circuit board PCB World Health Organisation WHO Occupational Safety Health and Administration OSHA Two-dimensional 2D Radially aligned nanorutile RANR TEM Transmission electron microscopy xxiii SEM Scanning electron microscopy PXRD Powdered X-ray diffraction FTIR Fourier transform infrared spectroscopy XPS X-ray photoelectron spectroscopy DRS Diffuse reflectance spectroscopy UV-Vis Ultraviolet-visible spectroscopy BET Branauer-Emmett-Teller (BET) HAL Hole accumulation layer BJH Barrtt Joyner Halenda 1 CHAPTER 1: INTRODUCTION 1.1 Background and motivation Volatile organic compounds (VOCs) can be simply defined as chemicals that have the ability to vapourize in ambient environment. These substances are omnipresent in both indoor and outdoor premises. The impact of indoor environmental contamination and pollution arising from VOCs has not been taken very seriously until very recently. The need for monitoring VOCs is very paramount since most people carry out their daily activities in enclosed premises. Statistics from World Health Organization (WHO) shows that close to 4 million of people die yearly due to indoor pollution yearly 1. Chemicals like acetone, ethanol and formaldehyde have been identified as common indoor pollutants. In a country like the United States of America, the environmental protection agency (EPA) regulates VOCs outdoor due to the possibility of formation of photochemical smog under precise conditions. According to WHO, organic pollutants are classified under three categories: very volatile organic compounds, volatile organic compounds, and semi volatile organic compounds. The classification is shown in Table 1.1. Table 1.1: Classification of volatile organic pollutants (adapted from WHO) 2. Description Abbreviation Boiling point range (⁰C) Examples of compounds Very volatile organic compounds VVOC <0 to 50-100 Propane, butane, methyl chloride Volatile organic compounds VOC 50-100 to 240-260 Formaldehyde, d- limonene, toluene, acetone, ethanol, isopropanol, hexanol Semi volatile organic compounds SVOC 240-260 to 380-400 Pesticides, plasticizers, and retardants 2 VOCs are generally available at home and work; it is inevitable not to be exposed to airborne VOCs. Common sources of VOCs include oil and gas field sites, industrial chemicals, fuel combustion such as petrol, wood, coal, natural gas, printing presses, pharmaceutical plants, and from solvents such as glues and paints. Industrial processes and emission from automobile vehicles have been reported as the main sources of outdoors VOCs. The era of nanotechnology has paved way for a rapid detection of volatile chemical compounds which have great advantages over the conventional methods of detection of chemical compounds such as gas chromatography (GC), ion mobility spectrometry (IMS) and reaction mass spectrometry. The development of nanotechnology has led to production of nanosensors such as breath analysers, E-nose and wearable nanosensors which provide a quick, affordable, harmless and a portable device for measuring VOCs with high sensitivity, good selectivity, ultra-fast response-recovery times and good repeatability. Chemical sensors find application in the following fields: healthcare-genetics, diagnostics, drug discovery, food processing, environmental and industrial monitoring, quality control, defence and security 1,3,4. Since the discovery of graphene, a two-dimensional (2D) material with great exotic properties and wide applications. But its application is not extended to optoelectronic applications due to lack of band gap. This have resulted into avalanche of other 2D nanomaterials among them are transition metal dichalcogenides (TMDCs) 5. Fascinating chemical and physical characteristics of 2D materials have made them to be explored in sundry applications such as thermoelectric device, field effect transistors (FET), photodetectors, solar cells, fibre lasers, optics, tribology, electrode materials, catalysis and sensors in the last two decades 5,6,7,8,9. TMDCs materials represent a new set of sophisticated class of materials that possess a layered structure just like the clay structure, it may have single or few atoms depending on the thickness of the nanomaterial 10,11,12,13. A TMDC has a structure which is quite different from graphene. A monolayer of TMDC is a three atoms stick which is made of a layer of transition metal atoms such as Zr, Hf, Ti, Ta, Mo sandwiched between two planes of chalcogen atoms such as S, Se and Te 13,14. These materials have some fantastic characteristics that they exhibit when they exist as mono or few layered materials. 2D nanomaterials have been deployed as sensors because of their layered structures. They possess high surface area and special semiconducting features arising from the manipulation of their band gap makes them suitable in sensor application 15,16. 3 Unlike the group VIB members such as MoS2, WS2, MoSe2 and WS2 which have been investigated widely, few works are available on group IVB despite theoretical projections showcased about these nanomaterials as possessing astonishing properties. Theoretical speculation has proved TiS2 has a good potential for chemical sensing. It is believed that TiS2 has more active sites for gas absorption than MoS2 owing to crystal structure and bonding that existed between the metal and the sulphur atoms 17. It was also predicted that monolayer ZrS2 could find applications in chemical sensing, however ZrS2 with a defect such as S-vacancies will behave as a better sensor compared to ZrS2 with no defect 18. Presently, the application of ZrS2 been demonstrated as an optical coupled plasmon waveguide resonance (CPWR) sensor but not as a chemirestive sensor 19. Decoration of polymers or semiconductor metal oxides with TMDCs could greatly enhance the sensor attributes arising from synergistic effects. Shokouh et al fabricated a nanohybrid of TiS2 and polyvinylpyrrolidone (PVP) for sensing ethanol vapour, the sensor displayed an ultra- fast response-recovery times of 2 s and 60 s respectively. The delayed recovery time that is often associated with TMDCs has been eradicated by forming close interjections between the nanocomposite 17. Qin et al made a nanohybrid of TiO2-WS2 by loading quantum dots of TiO2 on WS2 nanosheets for sensing NH3 gas. The sensor displayed better response (17 folds) and recovery times, selectivity and stability compared to bare WS2 20. 1.2 Problem statement The poor or no sensitivity of metal oxide-based semiconductor sensors to chemical vapours or gases at room temperature has prevented their usage in detection of flammable and explosive gases; this has caused a great shift to two-dimensional (2D) nanomaterials. Researchers are now investigating other 2D materials since graphene has properties that make it applicable in gas sensing. The layered structure and large surface-to-volume ratio of 2D nanomaterials make them suitable for gas sensing, with a fascinating good sensor characteristic towards volatile organic compounds. Therefore, these nanomaterials have been explored recently for their gas sensing attributes. However, they still have drawbacks like poor sensitivity and recoverability. Several studies have shown formation of hybrid nanomaterials could greatly improve the performance of transition metal dichalcogenides in gas sensing. The construction of proper close interface between the composite will enhance the electrical and chemical functionalities of the sensor. Integration of TMDCs with metal oxides, polymers and carbon materials will 4 boost the sensor characteristics of TMDC-based sensors with a geometric advantage of the nanocomposite material. 1.3 Aim and objectives The aim of this research is to synthesize nanocomposite catalysts of the transition metal dichalcogenide, ZrS2 with dandelion shaped TiO2 and polyaniline; and to evaluate their performance in the chemical sensing of volatile organic compounds at room temperature. To achieve the above-mentioned aims, the following were identified: 1. Colloidal synthesis of ZrS2 nanomaterials using hot- injection and heat-up methods. 2. Determine the optimal conditions for the synthesis of ZrS2 nanomaterials by varying the reaction time, coordinating solvents, temperature of reaction and concentration of precursors. 3. Fabrication of few layered ZrS2 from the bulk crystal sample via liquid exfoliation with green solvents. 4. Synthesis of dandelion shaped TiO2 by facile hydrothermal method. 5. Synthesis of polyaniline (PANI) using facile chemical polymerization method. 6. Functionalization of both the titania and polyaniline by decorating them with ZrS2 via ultrasonication. 7. Characterization of the synthesized nanomaterials and heterostructures using techniques such as SEM, TEM, HRTEM, UV-Vis, AFM, FTIR, BET, TGA, Raman spectroscopy, XRD and XPS. 8. Application of pristine and heterostructure nanomaterials in sensing of VOCs at room temperature. 1.4 Thesis outline Chapter 1: Introduction This chapter enumerates the details of this research in topics. It provides insight into the background, problem statement, motivation, aim and objectives. Chapter 2: Literature review 5 This chapter discusses the current state of research with respect to advances in the synthesis routes of group IVB transition metal dichalcogenides and their nanocomposites with metal oxide or polymer for chemical sensing of volatile organic compounds at room temperature. Chapter 3: Colloidal synthesis of ZrS2 The chapter is a journal article submitted to the Journal of Crystal Growth and is currently under review. The article reports on the novel synthesis of ZrS2 via the colloidal approach. The effects of using different sulphur precursors, different ligand and co-ligands were also investigated. The stability of ZrS2 was also probed. Chapter 4: Synthesis of rutile titania This chapter entails facile fabrication, characterization of radially aligned nanorutile titania and their applications for sensing primary alcohol and acetone gas. Chapter 5: Synthesis of ZrS2 nanomaterials This chapter demonstrates a facile top-down approach of obtaining few-layered ZrS2 via ultrasonication and its characterization. Chapter 6: Nanocomposite of PANI-ZrS2 and its applications in chemical sensing This chapter describes the fabrication of nanocomposite of ZrS2 and PANI, characterization and its application for sensing selected primary alcohol and acetone at room temperature. Chapter 7: Nanocomposite of TiO2-ZrS2 In this chapter, the synthesis of ZrS2 and TiO2, characterization and its sensing potential towards volatile organic compounds is outlined. Chapter 8: This chapter presents a general conclusion from the research, as well as recommendation for future work. 6 1.5 References 1. Conti, P.P., Andre, R.S., Mercante, L.A., Fugikawa-Santos, L. and Correa, D.S., 2021 Discriminative detection of volatile organic compounds using an electronic nose based on TiO2 hybrid nanostructures. Sensors Actuators, B Chem., 344, pp.130124. 2. McAughey, J.J., Pritchard, J.N. and Black, A., 1990. Risk assessment of exposure to indoor air pollutants. Environ Technol., 1990;11(4), pp.295-302. 3. Sharma, S. and Madou, Marc., 2012. A new approach to gas sensing with nanotechnology. Royal society, 370, 2448-2473. 4. Munawar, A., Ong, Y., Schirhagl, R., and Tahir M.A., 2019. Nanosensors for diagnosis with optical, electric and mechanical transducers. RSC Adv., 9, pp.6793- 6803. 5. Anichini C., Czepa, W., Pakulski, D., Aliprandi, A., Ciesielski, A., Samorì, P., 2018. Chemical sensing with 2D materials. Chem Soc Rev., 47(13), pp.4860-4908. 6. Pandit, A. and Hamad, B., 2021. The effect of finite-temperature and anharmonic lattice dynamics on the thermal conductivity of ZrS2 monolayer: Self-consistent phonon calculations. J Phys Condens Matter., 33 (42):1-23. 7. Mattinen, M., Popov, G. and Vehkamäki, M., 2019. Atomic Layer Deposition of Emerging 2D Semiconductors, HfS2 and ZrS2, for Optoelectronics. Chem Mater., 15, pp.5713-5724. 8. Wen, Y., Zhu, Y. and Zhang, S. Low temperature synthesis of ZrS2 nanoflakes and their catalytic activity. RSC Adv., 5(81), pp.66082-66085. 9. Li, L., Lv, R. and Wang., 2019. Optical nonlinearity of ZrS2 and applications in fiber laser. Nanomaterials., 9(3), pp.315. 10. Choi, W., Choudhary, N., Han, G.H,, Park, J., Akinwande, D. and Lee, Y.H., 2017. Recent development of two-dimensional transition metal dichalcogenides and their applications. Mater Today, 20, pp.116-129. 11. Tedstone, A. A., Lewis, D.J. and O’Brien, P., 2016. Synthesis, Properties, and Applications of Transition Metal-Doped Layered Transition Metal Dichalcogenides. Chem Mater., 28(7), pp.1965-1974. 7 12. Shi, Y., and Li, L.J., 2017. Synthesis of transition metal dichalcogenides. In: 2D Materials: Properties and Devices. Cambridge University Press, pp.344-358. 13. Yan, C., Gan, L., Zhou, X., et al., 2017. Space-Confined Chemical Vapor Deposition Synthesis of Ultrathin HfS, Flakes for Optoelectronic Application. Adv Funct Mater., 27(39), pp.1-9. 14. Singh, A.K., Kumar, P., Late, D.J., Kumar, A., Patel, S., and Singh., J., 2018. 2D layered transition metal dichalcogenides (MoS2): Synthesis, applications and theoretical aspects. Appl Mater Today., 13, pp.242-270. 15. Kumar, R., Goel, N., Hojamberdiev, M., and Kumar, M., 2020. Transition metal dichalcogenides-based flexible gas sensors. Sensors Actuators, A Phys. 303, pp.111875.-111892. 16. Lee, E., Yoon, Y.S. and Kim, D.J., 2018. Two-Dimensional Transition Metal Dichalcogenides and Metal Oxide Hybrids for Gas Sensing. ACS Sensors., 3(10), pp. 2045-2060. 17. Hosseini-Shokouh S.H., Fardindoost, S., and Zad, A.I., 2019. A High-Performance and Low-Cost Ethanol Vapor Sensor Based on a TiS2/PVP Composite. ChemistrySelect, 2 4 (21), pp.6662-6666. 18. Nguyen, H.T.T. Hoang, D.Q., Dao, T.P. Nguyen, C.V., Phuc, H.V., Hiew, N.N., Hoat, D.M. Long, H.L., Tong, H.D. and Pham, K.D., 2020. The characteristics of defective ZrS2 monolayers adsorbed various gases on S-vacancies: A first-principles study. Superlattices Microstruct.,140, pp.88-104. 19. Ma, J., Liu, K., Jiang, J., Xu, T., Wang, S. Chang, P., Zhang, Z.,Zhang, J., and Liu, T., 2020. All optic-fiber coupled plasmon waveguide resonance sensor using ZrS2 based dielectric layer.Optica, 28(8), pp.11280-11289. 20. Qin Z., Ouyang, C., and Zhang, J., 2017. 2D WS2 nanosheets with TiO2 quantum dots decoration for high-performance ammonia gas sensing at room temperature. Sensors Actuators, B Chem., 253, pp.1034-1042. 8 CHAPTER 2: LITERATURE REVIEW 2.1 Transition metal dichalcogenides for chemical sensing Among the new materials for sensor application, much attention is being given to two dimensional (2D) chalcogenide materials which have diverted a great research to explore them in the last few decades 1,2,3,4,5,6 . Transition metal dichalcogenides (TMDCs) materials represent a new set of sophisticated class of materials that possess a layered structure just like clay structure, it may have single or few atoms depending on the thickness of the nanomaterial 6,7,8,9. The shift to TMDCs emanated as a result of the exfoliation of graphene about two decades ago, a single-layered carbon with outstanding properties such as conductivity, mechanical, thermal, and electrical properties. Unlike graphene, a TMDC monolayer is three atoms thick consisting of a layer of transition metal atoms such as Zr, Hf, Ti, Ta, Mo sandwiched between two planes of a chalcogen atom such as S, Se and Te 9,10. There are weak van der Waals forces that exist in between the layers and covalent bonds within the materials. These materials have some fantastic characteristics that they exhibit when they exist as mono or few layered materials. 2D nanomaterials have been deployed as sensors as a result of their layered structures. They possess a high surface area and special semiconducting features arising from the manipulation of their band gap making them suitable in sensor application 2,3. TMDCs have some drawbacks which limit their application in gas sensing such as: (i) cross selectivity-the sensor may respond to many gases at a time, without being able to distinguishing one gas from the other (ii) TMDCs sensor response time is very slacking and the recovery time is often partial (iii) oxophilicity of TMDCs which refers to the tendency of the sensor surface to be somewhat topped by oxygen or moisture in ambient environment. This may slowly devalue the activity and stability of the sensor over time. Numerous methods have been used over the years to eradicate these drawbacks, which include novel metal doping and forming a junction with metal oxide semiconductors 3. In relation with TMDC materials, the monolayers and few-layers of hafnium dichalcogenides have astonishing physical properties speculatively, but are rarely investigated 9,12,13,14. Results have demonstrated that, the calculated room temperature mobility for HfS2 and HfSe2 monolayers is about 5.3 and 10.3 times respectively; greater than properly investigated MoS2 9,15. They also exhibited greater sheet current densities much higher than that of MoS2, this remarkable quality makes them valuable materials in FETs application 15. Notwithstanding these eminence in theoretical projection; correspondingly few experimental results are available presumably due 9 to the complexity in syntheses methods or as a result of oxophilicity of the nanomaterials 16,17. In the few published works, most of the studies use exfoliation method or chemical vapour deposition (CVD) method for their fabrication 14,18,19,20,21. The thickness and lateral dimension of the nanosheets cannot be manoeuvred with the use of the mechanical exfoliation method 15,18,19. CVD is one the favourite methods of producing the group IVB TMDCs except for TiS2 21,22. The method is economical, easy to carry out and produces a product with high quality; however, the CVD method still has some demerits. Ultrathin 2D TMDCs are mostly synthesized on substrates like mica, sapphire, SiO2 and hexagonal boron nitride (hBN) which require them to be conveyed to other substrates for further application 6,11,13,23. For application purposes, it is better that these nanosheets are in discrete form than to be produced on substrates 24. Hence, the solution method is still preferred for the synthesis of 2D TMDCs to CVD method. There are three common ways of preparing mono or few-layers TMDCs: (a) exfoliation, (b) chemical vapour deposition and (c) wet chemical/solution -based. 2.2. Exfoliation This method of synthesis can be subdivided into two - mechanical and liquid exfoliation. It is employed in the fabrication of mono and few-layers of 2D layered materials. Mechanical exfoliation is a physical process which requires the use of the scotch-tape technique to separate off from the bulk of the material. The adhesive force of the tape helps to break the van der Waals forces that exist within the layers of the material. The isolated layer can undergo successive peeling to give mono and few-layers sample which is then deposited onto a substrate. This produces high quality single-layered nanosheets though with low output while liquid exfoliation is a solution-based method with great output. The first report on exfoliation of TMDCs from bulk samples was published in 2005 by Novoselov et al where mono layers of MoS2 and NbSe2 were isolated from the bulk of the material using scotch-tape 25. Group IVB TMDCs have been prepared using mechanical exfoliation of the bulk. The group IVB TMDC nanomaterials are easily affected in ambient environment and thus get oxidized 26,27,28,29. To prevent their oxidation, exfoliation was carried out in a glove box and vacuum transfer chambers or immediate passivation with protective encapsulation layer was used in the synthesis 18,30,31. Chae et al fabricated few layers of HfS2 field effect transistor on a SiO2/Si substrate inside an integrated vacuum cluster system to prevent ambient oxidation. They realized uniform ambient oxidation of the HfS2 material, preferentially at defect sites which 10 resulted to thickness enlargement. The oxidized HfS2 performed poorly as a field effect transistor compared to the unoxidized sample 30. In a similar report Kanazawa et al synthesized few layer nanosheets from the bulk using scotch tape on single crystal HfS2, the small piece that was obtained on the tape, was then cleaved numerous times to obtain thin film of average size of thickness on SiO2/Si/Al2O3 substrate. The Al2O3 substrate surface was passivated with hexamethyldisilane (HMDS) before exfoliated HfS2 was transferred on it 32. In another similar work by the same group, HfS2 flakes were exfoliated on AlO3 substrate and the exfoliation method used suffered from the difficulty in manipulating the size and thickness of the fabricated TMDCs, as well as not being able to be scaled up for large batch production 26. Liquid exfoliation can be further subdivided into two main types. The one type is simple and does not involve intercalation. The bulk sample is dispersed into the appropriate solvent or surfactant followed by exfoliation through a sonication process. The second type is a two-step process with intercalation preceding the exfoliation process in a solvent 19,20,33,34. This method entails insertion of alkali metals into the bulk material with compounds like LiBH4, n-butyl lithium or organolithium compounds in solvent for 7-14 days at room temperature or at 100 ˚C for 3-4 days, succeeded by dispersing in an appropriate solvent. Care must be taken to ensure complete exfoliation, if not there is the tendency of generation of metal nanoparticles and Li2S being precipitated during the process. The lithiated layered material is recovered through a filtration technique and thorough rinsing with solvent such as hexane to eliminate lingering impurities of organic residues and the alkali metal. The intrinsic properties of the dispersing solvent or surfactants is crucial which is to break the cohesive energy that exist in the layered material and also determines the exfoliation output 33,35. The surface tension of the solvent needs to match the surface energy of the dissolved bulk sample. It is necessary to use the right solvent for dispersion of the bulk powder as this prevents re-stacking and aggregation of nanosheets in the solvent. Coleman et al reported that experimental parameters such as the starting mass of the bulk material, sonication period, centrifugation conditions and nature of solvent determines the concentration, thickness, lateral and broad size of the exfoliated nanosheets produced. An increase in concentration was favoured by increasing the sonication time (200 h) for MoS2, moreover, the nanomaterial had small lateral sizes and broad size distribution 20. The solvents that are commonly used are isopropanol (IP), dimethylformamide (DMF), hexane, N-methyl-cyclohexyl-2-pyrolidone (NCHP) and N-methyl-2-pyrrolidone (NMP) 35. NMP is the most suitable solvent for the fabrication of MoS2 but due to its toxicity and difficulty in obtaining free standing nanomaterial after sonication, its application is limited. 11 This has led to the use of other solvents such as aqueous solution or volatile solvents. Coleman et al used this method to prepare few-layers of some TMDCs and metal chalcogenides. For a solvent to be suitable for isolation of MoS2, it was observed its surface tension should be about 40 mJm-2 20. However, when water is used as a dispersing agent for a material that hydrolyses in water, the nanomaterial gets oxidized. Traces of water have a tendency to oxidize group IVB TMDCs through hydrolysis, resulting into the production of metal oxides (MOXs) 26,27,30,36. Sherrell et al gave an insight to the oxidation of TiS2. The group synthesized TiS2 nanosheets by insertion of alkali ions into TiS2 powder exfoliated in deionized water thereafter. It was observed that the exfoliated nanosheets were quickly destabilized by oxygen. The TiS2 nanosheet suspension in water was oxidized to oxide of titanium. It first generated TiSO species as the intermediate product, at the same time H2S gas was liberated to the environment. This was evident, as a colour change was observed in the suspension, it first turned to grey and then to white within 7 days of observation 36. This method is so popular in the synthesis of single layers of MoSe2, WS2, MoS2, WS2, TaS2, TiS2 and ZrS2 35. The obtained MoS2 and WS2 nanosheets using this method were observed to undergo phase transition, it translated from semiconducting (2H) phase to non-semiconducting (1T) phase. A thermal annealing process was needed to reverse it to the semiconducting phase 37. The use of co-solvents has been developed to enhance the exfoliation process. For this, the use of Hansen solubility parameters (HSP) theory must come into play. Using HSP theory, a variety of a mixture of solvents have been explored to fabricate MoS2. Zhang et al exfoliated MoS2 nanosheets using a co-solvent of water and ethanol. In a similar way, MoS2 was also effectively exfoliated using a mixture of chloroform and acetonitrile 38. Kaur et al synthesized few-layers of HfS2 by dispersing the bulk HfS2 in NCHP and ultrasonicated the solution. For the exfoliation in NCHP, the sheet formed was more stable in air as the solvent shielded the nanomaterial against ambient oxygen for a few days compared to using NMP and DMF as solvents 39. Li et al synthesized ZrS2 nanosheets by dispersing the bulk of the powder in IP, the suspension was then sonicated to give a few layers of ZrS2 27. Zeng’s group were the first to report on simplified lithiation using electrochemical lithium intercalation method to produce single layers of MoS2, WS2, TaS2 and TiS2 by proper adjustment of the amount of lithium intercalated, which was followed by exfoliation in ethanol or water 40. In another related experiment, the same group executed a systematic study by manipulating circuit parameters such as voltage and current for the for production of few- 12 layered inorganic compounds, such as TiS2, TaS2, WSe2, ZrS2, NbSe2, BiTe3, Sb2Se2 and BN. They optimized the parameters and produced high quality NbSe2 and BN nanosheets 40,41. Tandem molecular intercalation (TMI) is an improved method of exfoliation which is a facile singe step process that does not involve sonication, operated under safe and mild conditions. This method makes use of Lewis bases (short and long alkylamines or alkoxides) intercalates, in which short initiator molecules will be the first to intercalate, followed by long primary molecules. Group IV and V TMDCs are better synthesized using weak Lewis bases such as alkylamines while alkoxides are more appropriate for the synthesis of group IV TMDCs. Single-layered WSe2 has been exfoliated by intercalation of an alkali ethoxide and alkali hexanoate in dimethyl sulfoxide (DMSO) with agitation lasting for several hours at room temperature. This process is well accepted for the synthesis of TiS2, ZrS2, NbS2 and MoS2 42. The use of surfactants or polymers has not been well explored like the use of solvents presumably due to the high cost or toxicity of some of these surfactants. Sodium chlorate is the common surfactant that is being used, it helps to coat the sheets in dispersion thus preventing agglomeration 43. 2.3 Gas-Phase method of metal powder production Gas-phase method of synthesis is a bottom-up method of production of materials from unit atoms. Chemical vapour deposition and chemical vapour are gas-phase methods that are commonly being deployed in TMDC synthesis in particular group IVB TMDCs 17. 2.3.1 Chemical vapour deposition Chemical vapour deposition (CVD) is a methodology that involves decomposition or chemical reactions of gaseous precursors through thermally induced means. Depending on the nature of material or the application, sometimes the use of subtrate may be introduced into the furnace onto which the product is formed. CVD is generally deployed as a bottom–up method for synthesis of various 2D materials in the last decade, especially the group IVB TMDCs 6,11,13,44,45,46,47,48. In the synthesis, an inert gas (e.g., Ar) and H2 gas are introduced, which help to eliminate oxidation of the material, at the same time reduce formation of impurities 9,13,17,23,49,50. 13 2.3.2 Chemical vapour transport There are more reports on the use of chemical vapour transport (CVT) on the synthesis of group IVB TMDCs than mere CVD 11,51. CVT requires the use of halogens such as I2 as the transport gas and is used for bulk single crystal, which takes days for the synthesis to be completed. The synthesized materials are then exfoliated into single or few layered sheets 52. CVT is also used to synthesize transition metal trichalcogenides (MX3) which are then subjected to pyrolysis to give MX2 27,53. Wen et al synthesized ZrS2 nanoflakes by reacting a stoichiometric ratio of elemental sulphur (S) and ZrCl4; S was added in excess due to its ease of evaporation at high temperatures. Various temperatures were evaluated, and in each case, the reaction was held for 1 h. ZrS3 was formed and later decomposed to ZrS2 on further heating. The optimal temperature was observed at 800 ˚C with no traces of impurities. 50. In a similar manner with little adjustments, Fu et al synthesized HfS2 by using HfCl4 powder and S powder. Both the metal precursor and S powder were placed upstream of the quartz tubes and heated at a temperature of less than 200 ˚C while the substrate was inserted in the hot zone at a temperature of about 930 ˚C. The heating was done simultaneously for a few minutes and then stopped. The unreacted precursors were immediately eliminated with the aid of magnets, after which the furnace was cooled down naturally 13. Zheng et al similarly synthesized HfS2 nanoforests by placing the substrate in the hot zone; and the metal precursor and chalcogen precursors upstream. The heating was done at the same time, the temperature was set at 950 ˚C and 160 ˚C respectively, operated for 10 minutes and then terminated. At the end of the reaction, the furnace was rapidly cooled down 54. Zhang et al comparably reported on synthesis of both ZrS2 and ZrSe2 by CVD at elevated temperatures above 800 ˚C synthesis. The reaction temperature was sustained for about 20 minutes and nanostructures of ZrS2 and ZrSe2 were grown on substrates 23. Yan et al synthesized HfS2 nanoflakes using S powder and HfCl4 or HfO2 powder as precursors. The precursors were placed upstream and the substrate at downstream and the reaction was operated at 900 ˚C for just 10 minutes followed by cooling the furnace 9 . Shimazu et al synthesized a single crystal of ZrS2 by heating Zr, S8 and I2 in a sealed evacuated quartz operated for 3 days at 800 - 900 ˚C 11. TiS2 was also prepared by Gao et al using three temperature zones in which Ti/NH4Cl and S powder were placed at upstream where low temperature was applied, and the substrate was placed at downstream operated at 450 ˚C 55. CVD is not commonly used for titanium dichalcogenide synthesis, they are often produced through solid-state reaction method. A mixture of titanium powder and sulphur powder are blended together before being transferred and sealed into an ampoule under a 14 vacuum. It is then calcined using muffle furnace for 12 h at 500 ̊ C. Afterwards, the temperature is elevated to 800˚C and maintained for another 24 h after which the reaction is stopped and allowed to cool to room temperature 56,57,58. Fabrication of high quality 2D ultrathin TMDCs is very difficult as this method does not allow easy manipulation of reaction parameters. 2.4 Wet chemical synthesis This method involves the use of surfactants or polymers in the solution during the synthesis process. It is much easier for the fabrication of nanomaterials at low temperatures such as 130 ˚C and offers better control of kinetic parameters; unlike CVD which requires high temperatures of at least 400 ˚C and may run for several hours before the reaction gets to completion 59. The wet chemical synthesis method has four main variations which include colloidal, hydrothermal, sol gel and liquid exfoliation techniques. 2.4.1 Colloidal synthesis Colloidal synthesis provides favourable merits, such as easy to direct and proper grip of the crystallinity, monodispersity and control over the edges of TMDCs 60,61,62. The method has been deployed for the production of quantum dots, metal nanoparticles, nanomaterials, etc. When applied for the synthesis of TMDCs, reaction variables which include reaction time, temperature, nature of the metal precursor, chalcogen precursor and the type of ligand used are crucial in determining the shape and size of the nanomaterial that is formed 59,63. This synthesis technique can be further subdivided into two: hot injection and heat-up (one-pot synthesis) 59,64,65. Injection synthesis is often used when either or both the reactant(s) is/are solvent(s). The precursor that is in solid form is first dissolved in an organic solvent (ligand), then purged for about 20 minutes. The temperature of the system is raised at a controlled rate to a predetermined temperature, at this point the injection of the other precursor is introduced into the hot system by means of a syringe under vigorous stirring. The injection may be rapid or slow depending on the nature of the product required. One-pot synthesis is a non-injection technique which is commonly used when both metal and chalcogen precursors are solids. In this case the reactants are first mixed with the surfactant at room temperature before heating is introduced under inert conditions. Ligands that are commonly used in the colloidal synthesis of TMDCs include oleylamine (OLA), 1-hexadecylamine (HDA), oleic acid (OA), oleylalcohol (OYA), trioctylphosphine oxide (TOPO), dodecylamine (DDA), squalene, 1- dodecanethiol (1-DDT), stearic acid, octadecylamine (ODA) and 1-octadecene (ODE) 64,65,66. The use of ligands in colloidal synthesis helps in controlling the morphology during the 15 synthesis of semiconductor nano-/micro- crystals by coordinating to the surface of the growing nanoparticles 66. This method of synthesis is quite popular because of its simplicity. The use of ligands has eased the synthesis of hierarchical structure based morphologies such as comb– like, disk–like, dendrite-like, snow flake-like, rod-like, flower–like and urchin–like structures, which show unique properties by combining the features of micrometer and nanometer building blocks in one crystal 65,67. The colloidal method has been employed in the synthesis of single or few-layered TMDCs. Caution is needed in the synthesis of group IVB TMDCs due to the oxiphilicity nature of single or few-layered nanomaterials to prevent contamination of the nanomaterials with metal oxides. There is a need to strictly avoid water or oxygen in the preparation of precursor solution. The chlorides of group IVB (TiCl4, ZrCl4 and HfCl4) are commonly used in the synthesis of these nanomaterials. These metal precursors are very hygroscopic; thus, they get hydrolysed while weighing in ambient air. Hence, weighing should be done in a glove box. Prabakar et al reported on simple colloidal synthesis of hierarchical structures of TiS2 using hot injection method by dissolving elemental S in non-coordinating ODE and at 300 ˚C, TiCl4 was injected and heated up for 15 minutes. The metal precursor was injected into the system at a lower temperature of 150 ˚C to produce a flake-like structure unlike the flower-like structure that was observed at the higher temperature 68. In 2008, Park et al synthesized mono layers of nanodisks TiS2 using OLA in dried form; S powder was dissolved in OLA at 110 ˚C, the mixture was brought to room temperature. Thereafter, titanium was injected, and the temperature slowly raised to 215 ̊ C under argon gas, the reaction lasted for 12 h. The sample obtained was vacuum dried. On exposure of the nanomaterial to air, the colour changed from black to brown; however, when it was kept refrigerated under nitrogen, the colour remained unchanged. The thickness of the TiS2 nano-disk was about 0.6 nm and a lateral size of 50 nm. Increasing the concentration of the chalcogen atom with a rapid rate of temperature increase helped control the lateral size of the nano-disks. When the chalcogen concentration was increased by 100%, a decrease in lateral size was observed from 50 nm to 34 nm. At the same time, when the concentration of sulphur was increased by a double fold, there was reduction in the lateral size. However, the synthesized TiS2 single layer nano-disks were easily destabilised at room temperature, the S atom was displaced by oxygen atom in air. The authors were able to prove this using time independent energy dispersive spectroscopy (EDS) and powder XRD 69. In related work, Cheon’s group synthesized TiS2 by dissolving TiCl4 in OLA followed by injection of CS2 into the reaction mixture. The reaction was maintained for 15 min at 300 ˚C. 16 Increasing the concentrations of both metal and chalcogen precursors was so pivotal in controlling the lateral size of nanomaterial formed. When the concentrations were increased to 2.4 times, it produced TiS2 of lateral size of 40 nm. The group changed the S source to S powder; one pot synthesis was deployed to produce TiS2 nanocrystals. The authours preferred the use of CS2 over elemental S based on production of highly reactive radicals by S powder which resulted in structural degradation of the nanocrystals produced. Presumably, it is better to use a chalcogen that contains carbon for the synthesis of group IVB TMDCs as the carbon can react with any traces of metal oxides formed during synthesis. The method was also employed in the synthesis of TiSe2 by using elemental Se as source of chalcogen. Not much has been reported in open literature on group IVB, in particular on hafnium dichalcogenides. Few reports are available on colloidal synthesis of HfS2. Cheon’s group reported on its synthesis, in which the reactants HfCl4, 1-DDT and OLA were heated up in a reaction flask in an inert atmosphere at 245 ˚C for the duration of 10 h 70. In a similar work, the group also used CS2 as a S precursor at higher temperature of 320 ̊ C 60. The same group also used a resembling synthesis procedure to synthesize sulphides and selenides nanosheets of the IVB and VB groups. To date, the studies by the Cheon’s group are the only ones available on ZrS2 synthesis 70,71,72,73. They distilled and degassed OLA prior to use in order to purify it. However, they did not mention the oxiphilicity and the ease of oxidation of group IVB TMDCs. In 2014, Cheon’s group demonstrated the use of slow decomposing chalcogen precursors (1-DDT) and H2S gas in the generation of single–layers nanosheets of Group IVB metal sulphides. At high temperatures (over 150 ˚C) 1-DDT releases S atoms gradually. In the synthesis of HfS2 nanosheet a combination of HfCl4, OLA and 1-DDT were heated up in a reaction flask in an inert atmosphere at 245 ˚C for 10 h. The 1-DDT slowly decomposed to H2S during the entire synthesis time. Using the same procedure while differing the temperature and time of synthesis, ZrS2 and TiS2 nanostructures were also produced. The group also used H2S gas in a similar set- up; the H2S was released slowly over 6 h and monolayers of TiS2, ZrS2 and HfS2 were obtained. A high level of precaution needs to be exercised when using H2S due to its toxicity which can be fatal. The nanocrystals obtained were of a poorer quality compared with those obtained with the use of 1-DDT 74. Different S precursors can be used, such as elemental S, 1-DDT, CS2, diphenylurea, thiourea and thioacetamide. CS2 and 1-DDT can be systematically introduced into the heating system using hot injection protocol while other precursors that are in solid form can be dissolved in the solvent prior to heating up with the metal precursor. CS2 is often used for hot injection as a result of its in-situ hydrogen sulphide generation. A metal chloride is 17 commonly used as the metal precursor and there is a need to wash the synthesized nanomaterials thoroughly due to lingering impurities such as chloride ions. Similar to group IVB, not much work has been reported for group VB. Sekar et al was the first to report on synthesis of group VB TMDCs. The group synthesized NbSe2 nanowires using one- pot synthesis. Into a three-neck reaction flask, a mixture of OLA, DDA and the precursors (NbCl4 and Se) were introduced. Temperature of the reaction was increased to 280 ˚C under N2 atmosphere and sustained for 4 h. The reaction vessel was not cooled down and nanomaterials obtained were rinsed with hexane. It was then subjected to heating under N2 atmosphere at 450 ˚C for a period of 3 h. The resulting NbSe2 wires had diameters varying from 2 - 25 nm. When the reaction vessel was cooled to room temperature before washing, nanoplates of NbSe2 were obtained. The nanoplates had lateral dimensions that range from 500-1000 nm and thickness between 10-70 nm. The authors also varied the reaction parameters in order to achieve different sizes or morphologies, but no changes were observed 75. Mansouri and Semagina reported on the synthesis of NbS2 using a mixture of ligands to produce different morphologies varying from nanosheets, nanospheres, nanohexagons and nanorods. Increased amounts of OA in OLA led to more production of the nanosheets, but better morphology and laterally confined 2D nanostructures were obtained with minimal use of OA. Also, of great importance is the timing of the reaction, with 0.25 h of reaction, monolayer of NbS2 nanosheets were formed with OLA as the sole solvent and as the reaction time was increased few-layers of nanosheets were formed 61. Han et al synthesized NbS2 by a one- pot synthesis method in which stoichiometric ratios of NbCl5 and 1-DDT were introduced into a three-neck reaction flask containing OLA in a glove box. The use of glove box indicated that the nanomaterial is sensitive to impurities (O2/H2O). The reaction was operated for 30 minutes at an elevated temperature of 280 ˚C after which it was cooled down. The authors do not mention if the reaction was conducted in an inert atmosphere. The synthesized ultrathin triangular NbS2 nanosheets had a thickness of 3.9 nm which was ascribed to represent five layers 37. A great number of research has been reported on group VIB TMDCs most especially on MoS2 and WS2 which have been demonstrated in sensor application. Mahler et al synthesized both prismatic 2H-WS2 and distorted octahedral 1T-WS2 structures 76. Hot injection reaction method was deployed, the metal precursor was added to OA in a vial, and the injection was made into a reaction flask containing OLA at 320 ˚C dropwise for about 0.5 h. Prior to attaining this temperature, CS2 was injected into the system. This resulted in controlled monolayers by the slow release of the precursors. The addition of HMDS after degassing helped in tuning the 18 crystal structure of the nanosheets from prismatic to distorted octahedral structure which was flowerlike in shape. In a similar report, Geisenhoff et al synthesized WSe2 using the hot injection method whereby tungsten hexacarbonyl was dissolved in combination with TOPO and OA and selenide precursor 77. The mixture was injected at 330 ˚C and the process was completed in half an hour. The ligand mixture helped to adjust the precursor reactivity and an increased amount of OA limited the metal precursor reactivity, resulting in fewer nucleation and thus bigger nanocrystals were formed. Lin et al prepared MoS2 quantum dots by dissolving the single-source precursors, ammonium tetrathiomolybdate ((NH4)2MoS4) in three different capping agents, OLA, OA and ODE 78. The reaction mixture was purged at 120 ˚C under vacuum for 2.5 h with stirring. Afterwards the reaction was sustained for 3 h at a temperature of 250 ˚C under N2 atmosphere before the reaction was quenched. In a related experiment, Altavilla et al prepared both MoS2 and WS2 nanosheets using thio-salts of Mo and W by a one- pot synthesis method. The single-source precursor and OLA was first degassed, and then the temperature was raised to 360 ˚C under N2 flow for 0.5 to 15 h. Interestingly, as the reaction time increased so did the number of layers produced 78. Antunez et al, in a similar experiment produced WSe2 nanosheets by injecting di-tertbutyl diselenide (tBu2Se2) into a reaction vessel already containing WCl4 in DDA at a temperature above 100 ˚C under N2 gas. The reaction lasted for 6 h under strong magnetic stirring at 225 ˚C. WCl4 is not a suitable metal precursor for the synthesis of group VI TMDCs, due to its ease to hydrolyse in air, thus the sample was introduced into a three-neck round-bottom flask in a glove box. It is well established that WCl4 is easily reduced in the presence of organics if overheated, hence the temperature during heating must be well controlled. The organic solvent (DDA) was not used as supplied but was deoxygenated and distilled before use 79. Jung et al, used a one- pot method to synthesize WSe2 monolayer nanosheets. W(CO)6, diphenyl diselenide were dissolved in OA, the system was degassed, and the temperature was raised to 330 ˚C, the reaction was operated for 12 h. Monolayer nanosheets with a lateral dimension of several nanometres were produced. The group also worked with other surfactants such as OYA and OLA. With OYA, few-layers (2-3) with lateral size of few nanometers were obtained while with OLA multi-layers with smaller lateral size were formed 80 . Zhou et al working with mixed surfactants, explored the influence of mixed solvents to produce different layers of the nanostructures of MoS2 and WS2 using injection protocols. The authors used Mo(AC)2 and W(CO)6, the surfactants stearic acid, TOPO, ODA and squalene under N2 atmosphere with stirring. The duration of heating was just 1 h before being quenched. In a similar reaction, MoS2 monolayers were also generated but with a change of surfactants (OA, stearic acid and TOPO). By increasing the concentration of 19 the chalcogen source and decreasing the concentration of the metal as well as increasing the temperature of the reaction multi-layers (3-5) of WS2, nanosheets were formed. With little modification, both thioacetamide and Mo precursor were injected into the reaction system at different times to produce few and multi-layers MoS2 nanosheets 81. The mechanism for the formation of single-layered nanosheets of TMDCs is still complex. Three approaches variables have been employed to optimise their formation: firstly, the use of chalcogen precursor that gradually decomposes over a long-time during synthesis or delay injection of reactant for a long period, secondly, the nature of the organic solvent; and lastly the duration of the reaction have been varied. 2.4.2 Hydrothermal or solvothermal synthesis Hydrothermal or solvothermal method is a versatile and effective synthetic route to produce nanomaterials with a different array of morphologies. Hydrothermal synthesis is one of the most important methods for producing fine powders of metal oxides. The process entails the reactants being dissolved in a solvent, which is then introduced into an autoclave. If the solvent for the reaction is non-aqueous, it is referred as solvothermal; whereas, if the solvent for the reaction media is water, it is termed hydrothermal 82. Teflon-lined autoclaves are used in this process; they are preferred over glass and quartz autoclaves, since they can tolerate high temperatures and pressures. Furthermore, they support alkaline solutions as well as being resistant to hydrofluoric acid. The flexibility of the method makes it easier to manipulate reaction parameters to produce nanomaterials with desired properties and quality. It has been extensively used for preparing metal oxide nanoparticles, chalcogenide, and phosphide nanomaterials 83,84,85. Chen and Fan synthesized NiSe2, NiS2, CoS2, FeS2, MoSe2 and MoS2 using hydrothermal synthesis at a low temperature while varying the synthesis parameters. They found that adjusting the reaction variables could extend the method of synthesis to other products 83. The method is commonly used for the synthesis of group VIB TMDCs and their nano composites 86,87,88,89,90. Huang et al synthesized MoS2 nanosheets with a net-like morphology of well linked nanoflakes via the hydrothermal route; the nanoflower material had a thickness of a few nanometres. The prepared material had a large surface area and good conductivity giving it potential for application in high-performance supercapacitors 88. Jang et al also prepared CdS nano particles using hydrothermal method as well as CdS nano wire/TiO2 nano particle composite using solvothermal method to produce effective photocatalysts 91. 20 2.4.3 Sol-gel method The sol-gel method is a popular avenue to fabricate metal oxide catalysts such as oxides of Ti and Si composites. A variety of materials which includes nanostructures, nanomaterial nanoparticles, glass, ceramics, and nanocomposite are generally fabricated using this method. This process generally takes place in three steps viz: hydrolysis, condensation, and drying. Sol- gel can be sub-divided into two types: aqueous sol-gel and non-aqueous sol-gel method 82,92. To synthesizie these colloids, the common precursors are made of metal alkoxides and alkoxysilanes. The use of tetramethyoxysilane (TMOS), and tetraethylorthosilicate (TEOS) is most common. There is a need to first make a homogeneous solution of the alkoxides to be used92. This method is very suitable for production of group VIB TMDCs and their nanocomposites 93,94,95,96,97,97. The method offers some merits such as low synthesis temperatures, high reproducibility, cost effectiveness and products with high purity, porosity, and large surface 98. A one-step direct sol-gel synthesis method was employed to prepare p- type few-layer MoS2 films in a large volume via deployment of Mo-containing sol-gel including 1% tungsten 94. It is very functional and good for large production using spin coating deposition method on a variety of substrates to produce 2H-MoS2 thin film having uniform surface area at moderate temperatures (300-400 ˚C) followed by annealing. The thin films MoS2 produced was of good quality and has great electronic properties with a narrow energy band gap of 1.35 eV which is consistent with the material. The product is of n-type semiconductor which finds application in electronic devices 95. 2.5 Gas sensing mechanism in TMDCs Semiconductor metal oxide sensors are popularly known for sensing of gases in the last decades especially toxic and dangerous gases, however the need of elevated temperature for their operation has inhibited their application in detection of flammable and explosives gases. The discovery of graphene, a 2D material for gas sensing has led to the emergence of TMDCs for gas sensing. TMDCs provide good selectivity, high sensitivity at room temperature due to ultrathin nature, high surface areas and adjustable bandgap which can find application in the detection of these gases. The key principle in gas sensing in TMDCs is principally based on charge-transfer activity connecting gas molecules and the immediate environment of the sensor which leads to a change in the electrical conductivity of the sensor. Electrostatic force is responsible for adsorption of gas molecules upon injection of the analyte gas on the 2D material. When air is flushed in the test chamber, desorption of the gas molecules takes place; 21 in this case there is restoration of the sensor to its original state. Figure 2.1 shows the mechanism of gas sensing of a typical TMDC with NH3 gas. The kind of interaction that exists between a gas analyte and sensing material will decide the level of the resistance adjustment. Based on the charge carrier of a semiconductor, it can be categorized as p-type or n-type. With a p-type semiconductor there is an increase in the resistance of the sensor upon interaction with a reducing gas whereas there is reduction in the resistance of the sensor with n-type semiconductor. The charge carriers on a p-type semiconductor are principally holes while it is electrons with n-type semiconductor. Due to high electron mobilities, TMDCs are proficient in sensing at room temperature and do not suffer from constant drift of signal which is typical of metal oxide occasioned by forwards and backwards distortion of oxygen vacancies since there is no oxygen atom present in the matrix of the sensor 2,3,5. Fig 2.1: Schematic illustration of sensing mechanism of TMDCs. The illustration is adapted from Gqoba et a l 99. 2.6 Attributes of a gas sensor Several parameters are very important in determining the performance of a sensor which include response and recovery time, the 3S (sensitivity, selectivity, and stability of the sensor), operating temperature, concentration of analyte, humidity, and detection limit. 2.6.1 Response The response is obtained by calculating the ratio of difference in conductance change of the analyte and in air and conductance in air. Mathematically this can be expressed as e- e- 22 R= 𝑅𝑔−𝑅𝑎 𝑅𝑎 = 𝛥𝑅 𝑅𝑎 (2.1) R= 𝑅𝑔 𝑅𝑎 (2.2) R= 𝑅𝑔−𝑅𝑎 𝑅𝑎 x 100 = 𝛥𝑅 𝑅𝑎 x 100 (2.3) R= 𝐼𝑜−𝐼𝑔 𝐼𝑔 x 100 = 𝛥𝐼 𝐼𝑔 x 100 (2.4) Reports have shown that the response and humidity have direct correlation. At extremely high humidity level, water vapour is adsorbed on the surface of the sensor by physisorption which may affect the electronic and ionic conducting properties of the sensor. The adsorbed water layer could serve as a barricade between the sensor surface and the analyte. Thermal treatment can be deployed to eliminate the humidity; however, the heating effect may result into agglomeration of the nanostructures. Reports have shown that acetone sensors perform poorly in terms of response at high temperature arising from thermal ripening of the sensor 100. 2.6.2 Response and Recovery Times The response time can be defined as the time it a takes a sensor to attain 90% saturation upon injection of analyte gas. The recovery time is the time taken to bounce back to 90% of no load upon withdrawal of the analyte gas and consequently blowing dry air to the sensor. A good sensor is expected to have rapid response and recovery times 100. 2.6.3 Selectivity / Specificity The pragmatic use of a sensor lies on this key parameter, the selectivity of a sensor estimates the preferential ability of a sensor to detect analyte gas of interest in the presence of unwanted gases. The value ranges from 0 to 1, the closer the value to 1 is an indication of high selectivity of the sensor 99. 2.6.4 Detection limits This can be defined as minimum concentration that can generate a signal response from a sensor. Two methods are commonly deployed in determination of LOD. Firstly, this is the most widely used method in which repetitive measurements are conducted without any introduction of any analyte gas and with introduction of minimum concentration of the analyte gas. 23 LOD = LOB + SD low concentration of analyte (2.5) Where LOD = Limit of detection LOB = Limit of blank SD = Standard deviation The second method entails obtaining LOD by extrapolation from a linear calibration curve 101. A good sensor should be able to detect a very low concentration of gas analyte. This is often reported as part per million (ppm) or part per billion (ppb). 2.6.5 Stability This is another key factor which determines the reliability and dependability of a sensor. It indicates the effectiveness and consistency of the sensor over a period. A good sensor should have a good stability close to 2-3 years and be recyclable. 2.6.6 Sensitivity and grain size The size of sensing material plays an important role in its sensing capability. A reduction in the grain size contributes to considerable enlargement in surface area and hence more surface sites for adsorption of analyte to react. This will therefore boost the sensing properties of the sensor 102. No model has been proposed for TMDCs however, three different models have been developed for a metal oxide sensor. Firstly, if the grain size (D) also known as crystallite size is much bigger than space-charge layer (L) D >> L, the large grain does not contribute to improved sensitivity. The electrical conductivity is governed by grain boundaries which control the inner mobile charge carrier and the gas mechanism. The second model suggested that when the grain size reduces, the L moves closer into each of the grain. When grain size is very close to but still higher than 2L, the depletion region around each neck produces conduction channels as shown in Figure 2.2. Both the cross-sectional area and grain boundary barriers determine the electrical conductivity. The sensitivity of the gas sensor depends on the grain size when D ≥ 2L. Thirdly, when the grain size undergoes reduction to 2 folds then D < 2L, the entire grain is depleted by the electrons, hence there is reduced conductivity as shown in Figure 2c. According to Dmorch et al the third model (D < 2L) is the most suitable system to improve gas sensing capability. Seal and Shukla reported using SnO2 as a case study; when the grain size is less than ⁓10 nm, this leads to better sensitivity for SnO2 sensor 102. Henceforth, a nanoparticle material is expected to display higher sensitivity compared to coarse-grained peer. Both the 24 receptor and transducer functions are influenced by grains. Therefore when the crystallite decreases, the higher the sensitivity of the sensor 103,100. Figure 2.2: Schematic highlighting the importance of grain size. Adapted from Dmorsh et al 1 00. 2.7 Factors controlling sensitivity Three important separate characteristics determines the response of a chemical sensor which are receptor, transducer, and utility 104. 2.7.1 Receptor This entails how individual grains of a sensor interact with immediate atmosphere accommodating oxygen and target gas, this influences its rate, selectivity, and reversibility. Adsorption only occurs as a result of high reactivity of the sensor surface. The absorption can either be by physisorption and chemisorption 104,105. 2.7.2 Tranducer 25 A transducer is a device that processes energy from one form to a readable signal. A tranducer function shows the capability of the sensor to translate the signal arising the interactivity occurring on the sensor surface into an electrical signal. A sensing material possesses a receptor and transducer functions, the boundary that exists within the grains plays this role of transducer function 104. 2.7.3 Utility factor This is related to the diminishing of the response of a sensor arising from factors of diffusion and penetration of the target gas across the pores of the bulk. This explains how the target gas migrates into the inner grains of the sensor material. When the reaction rate is higher compared to diffusion this leads to decrease in response, this results from inability of the target gas to reach inner sites of the grains of the sensor material 104. Figure 2.3 illustrates the relationship that exists among these three factors. Figure 2.3: Three factors determining the response of semicomductor gas sensor. Adapted from Yamazoe and Shimande 104. Utility factor (assembly issue)Transducer function (interparticle issue) Receptor function (interparticle issue) O- O- O-O- Space charge layer eH 2 O H 2 Double Schottky barrier (?) Gas diffusion and reaction 26 2.8 Gas sensing properties of 2D TMDCs and other 2D materials The group VI TMDCs are recently being applied in chemical sensing, they have the advantage of being able to respond to chemical vapour at room temperature. More studies have been reported on MoS2 compared to WS2. MoS2 has been demonstrated to sense NO2, NO, O2, NH3 and H2 at room temperature. Most of the early studies on the use of MoS2 as chemo-resistive gas sensor were conducted above room temperature. He et al reported on how adjustment of the number of layers can play a key factor in their sensing capabilities. It was observed that the thickness (layer) of the TMDCs plays a key role in determining the response to NO2 gas. The response of the sensor decreases on exposure to the gas with increasing thickness of MoS2, this is because of decrease in surface -to-volume ratio of the MoS2. The authors also reported better response of MoS2 over rGO based sensor. Integration of Pt nanoparticles with MoS2 nanosheets improved the sensitivity by a multiple of three 105 Late et al provided a contrary view in their studies where the five layered MoS2 activity was better than fewer layers (2-4). The insight of the mechanism responsible was not properly expantiated; it was merely attributed to electronic structures 106. Cho et al vertically grew MoS2 on SiO2 substrate using the CVD method. Three different arrangements were fabricated; vertical, horizontal, and mixed alignments. The vertical orientation’s performance was better than other alignments; this is ascribed to both the basal plane and edge sites providing more adsorption sites thus, better response 107. Kim et al produced microflakes and nanoparticles of MoS2 using mechanical and liquid exfoliation respectively. The nanoparticles, due to larger surface area and more edge sites than microflakes, produced better response to various O2 concentrations. The n-type behaviour of the MoS2 sensor is akin to metal oxide semiconductor in which pre-adsorbed O2 molecules is very paramount in the sensing mechanism. The sensitivity of the MoS2 sensor (liquid) for ethanol was high because of enormous physisorption of O2 molecules on the MoS2 108. Qin et al exfoliated bulk WS2 into few layers; the number of the layers plays a key role in the determination of the sensitivity of the sensor. With high number of layers, the response increases but poor recovery. This delay in recovery was expantiated using first principle calculations. With increasing number of layers, the binding energy that exists between the gas molecules and the layer(s) increases hence it leads to longer recovery times (desorption of gas molecules) 110. Cha et al functionalized surface edges of WS2 using multiple tubular carbon 27 nano-fibres (WS2@MTCNFs) which resulted into greater specific surface area. The responses increase with increasing concentration of the gas molecules. The sensor sensitivity was quite good toward NO2 than both NH3 and C7H8 111 . Several publications have shown WS2 sensors are very sensitive to NH3 vapour: Gqoba et al synthesized hierarchical nanoflowers of WS2. The selectivity of the sensor toward increasing concentration of NH3 vapours was good. The response of the sensor towards acetone, ethanol, toluene, and chloroform was quite rapid as well, the hierarchical nanoflowers’ morphology provided large surface area for the adsorption of the gas molecule. However, the sensor displayed poor recovery due to delayed desorption of the NH3 molecules after strong adsorption on the sensor 99. Late et al demonstrated single layer MoSe2 for NH3 sensing at room temperature with quick response (150 s) with increasing concentration of NH3 vapours but poor recovery (540 s). Despite poor recovery from the sensor, studies showed the performance is far better than some reported work on graphene and MoS2 sensors for sensing NH3 and NO2 at room temperature 112. In a related experiment, Guo et al exfoliated few layers WSe2; the fabricated sensor exhibited fantabulous sensitivity as low as detecting 0.05 ppm of NO2 gas with a rapid response time and recovery 113. SnS2 is a marvellous 2D material which is not part of the TMDCs, however it has a hexagonal structure akin to TMDCs with a layered structure. The material has a sizeable band gap when in few layers and large electronegativity that seems suitable for gas sensing application. In 2006 Shi et al was the first to demonstrate flower-like SnS2 synthesized using the hydrothermal process; the sensor had good sensitivity and selectivity for NH3 gas at room temperature with an average response and recovery time of 45 and 110 s respectively 114. Xion et al produced SnS2 nanoflowers using the solvothermal method. The hierarchical three-dimensional nanoflowers on exposure to 100 ppm NH3 produced optimum sensitivity at 150 ˚C, though with poor response/recovery time, however optimum response/recovery time was gotten at 200 ˚C. Oxygen catalysed the performance of the sensor; with increasing oxygen content (0-40%) on the sensor, this improved the performance of the sensor in about 3.5 folds. The DFT calculations were used to corroborate on the effect of the increasing oxygen content on the sensor performance. The sensor was able to determine concentrations of NH3 as low as 0.5ppm and the best selectivity was toward NH3 over CO2, CH4, H2, C2H5OH and CH3COCH3 115 . Wang et al demonstrated single layer SnS2 obtained via intercalation liquid exfoliation. The monolayer SnS2 sensor displayed a quick response (16 s) and good sensitivity towards 28 ammonia vapour. The response time was excellent, which is quite better compared to other NH3 gas sensors reported so far however, the author did not report on the recovery time 116. Guidi et al were able to demonstrate the of use of CdS and SnS2 and their metal oxide counterparts for sensing alcohols, acetone, and other chemicals. The CdS and SnS2 sensors were operated at room temperature with the aid of UV light as well as optimum temperature of 300 ˚C. The performance of the metal sulphide and SnS2 sensors outweigh the metal oxide sensors, the CdS sensor displayed great selectivity for alcohols while no selectivity from CdO. The highest selectivity was to butanol, followed by ethanol and methanol was the least. The SnS2 showed good selectivity to acetone vapour. The operating temperature was maintained at 300 ˚C. If exceeded, the S atoms and sulphur vacancies in SnS2 will be displaced with oxygen atoms 117. Only a few reports are available on chemical sensing of the novel nanomaterials of group IVB TMDCs. Shokouh et al reported a fascinating recovery with an ultra-fast response time of 2 s and recovery time of 60 s by fabricating nanohybrids of TiS2 and polyvinylpolymer (PVP) for ethanol sensing 1. Table 2.1 shows some of the TMDCs that have been demonstrated for chemical sensing. Table 2.1: TMDCs, metal oxides and their nanocomposites for chemical sensing Material T ˚C Gas (ppm) Sensitivity (%) Response time (s) Recovery time (s) References WS2@MTCNFs 25 NO2 15 - - 111 WS2 (single layer) 25 NH3 - 200 271.5 110 MoS2 400 O2 - 5700 - 109 MoS2 25 NO2 (0.1) - - - 108 MoS2 (Five layers) 25 NO2/NH3 1372/86 - - 107 MoS2 (single layer 25 NO2 (1.2) - >1800 >1800 106 29 WS2 25 NH3 (240) - 28 42 99 MoSe2 25 NH3 - 150 540 112 WSe2 25 NO2 (0.05) 5.06 50 105 113 SnS2 25 NH3 (5) 21.6 40-50 100-120 114 SnS2 200 NH3 (100) 7.4 40.6 624 115 SnS2 25 NH3 (50) 2.4 - - 116 SnS2 and CdS 300 CH3COCH3 and ROH - - - 117 Ag-TiO2 25 C2H5OH (5) 4.35 52 63 118 GaN-SnO2 25 ROH - ⁓100 ⁓100 119 Chitosan 25 C2H5OH 19.35 - - 120 WO3 300 C2H5OH 70 <15 <15 121 NiO-Sn 340 C2H5OH 15.6 17 15 122 MoS2-SnO2 25 NO2 (5) 18.70 - - 4 MoS2-SnO2 280 C2H5OH - - - 123 MoS2-ZnO 25 NH3 46.2 - - 124 MoS2-TiO2 150 C2H5OH 14.20 - - 125 WS2-TiO2 25 NH3 43.72 - 174.43 126 TaS2-PANI 25 H2O 97.00 36 49 127 TiS2-PVP 25 C2H5OH(2) 6800 60 2 1 30 2.8.1 Metal doping Transition metals which include Fe, Cr, Al, Ni, Sn, and Ce and precious metals such as Pt, Pd and Au can help boost the sensing properties of a semiconductor by enhancing the activity and adjusting the resistance of the sensing material. This improves the overall attributes of the sensor such as 3S (sensitivity, selectivity and stability), response and recovery time; and working temperature especially metal oxide sensors 103. Extensive reports have shown that doped ZnO sensors are more effective in sensing acetone vapour over undoped ZnO. Kaur et al in their review also showed doped 1- dimensional (D) titanium dioxide with precious metals are very effective in sensing ethanol and acetone compared to undoped titania 99. In a similar report, doped SnO2 sensor showed enhanced sensing capability to VOCs especially acetone unlike the undoped SnO2. More studies are yet to be available on functionalization of TMDCs with metals and precious metals; however, 2-D materials like graphene sensors doped with precious metals showed better sensing performance over pristine graphene 99. 2.8.2 Metal oxides for sensing VOCs Vast information is available on the sensing mechanism of the metal