WEARABLE GAS SENSOR MICROSYSTEM FOR PERSONAL HEALTHCARE AND ENVIRONMENTAL MONITORING By Xiaoyi Mu A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Electrical Engineering – Doctor of Philosophy 2013 ABSTRACT WEARABLE GAS SENSOR MICROSYSTEM FOR PERSONAL HEALTHCARE AND ENVIRONMENTAL MONITORING By Xiaoyi Mu Exposure to air pollutants is an ever present concern for human health in modern society. Most existing tools for monitoring air pollution are stationary and incapable of accurately measuring individual exposures that vary with personal lifestyles and environments. To study the health impacts of personal exposure, researchers need new monitoring tools that are suitable for long term personal use. This dissertation seeks to overcome the challenges of implementing an inexpensive, highly miniaturized, and rapidly responding gas sensor array system to measure air pollutants for personal health. By thoroughly studying the requirements of wearable gas sensing and the characteristics of prevailing gas sensing technologies, room-temperature-ionic-liquid (RTIL) electrochemical gas sensing technology is identified for the wearable microsystem development. To overcome the RTIL-based sensor’s inherent slow response, a planar porous-electrode-on-permeable-membrane (PEoPM) structure is introduced, demonstrating a 180% faster response to SO2 compared to a traditional structure. Microfabrication processes are utilized to miniaturize PEoPM sensors and construct a physically flexible sensing platform. To achieve a low-cost low-power sensory system, a new electrochemical instrumentation circuit topology is introduced. The new circuit uniquely utilizes the inherent nature of electrochemical sensor properties to significantly reduce circuit complexity compared to traditional architectures. Finally, this thesis presents the first ever single-chip CMOS electrochemical gas sensor that further miniaturizes the sensory microsystem. The monolithic CMOS-RTIL gas sensor occupies 2 0.48mm per sensing channel and demonstrates a limit of detection that is seven times better than a hybrid (multi-chip) solution. The results of this research lay a solid foundation for a personally wearable environmental monitor that could greatly improve human healthcare. ACKNOWLEDGEMENTS Though only my name appears on the cover of this dissertation, a great many people have contributed to its production. I owe my gratitude to all those people who have made this dissertation possible and because of whom my graduate experience has been one that I will cherish forever. First of all, I would like to thank my advisor Professor Andrew J. Mason for his continued invaluable mentorship, encouragement and inspiration through my PhD research. I would also like to acknowledge my dissertation committee including Professor Wen Li, Professor Prem Chahal and Professor Scott Calabrese Barton for their valuable feedback that made this dissertation possible. I must also thank my colleagues at AMSaC lab who helped me put my ideas to practice. Special thanks to Yue’s direction, idea sharing and answering my questions. I am grateful to Xiaowen, Haitao, Lin and Yuning for their help. It is my great pleasure to know and work with them. I would like to thank our collaborators Professor Zeng’s group. Special thanks to Dr. Wang’s valuable discussions and idea sharing. I also appreciate Professor Zeng and Min Guo for their cooperation, valuable knowledge share and discussions. Finally and most importantly, I want to thank my parents and brother for believing in me, and my wife for supporting me every step of the way. iii TABLE OF CONTENTS LIST OF TABLES ....................................................................................................................... viii LIST OF FIGURES ....................................................................................................................... ix 1 Introduction ....................................................................................................................... 1 1.1 Motivation .......................................................................................................................... 1 1.1.1 Air pollutants effect on human health and research method ........................................ 1 1.1.2 Air pollutant monitoring and data acquisition platform............................................... 2 1.2 Challenge ............................................................................................................................ 3 1.3 Goal .................................................................................................................................... 6 1.4 Thesis outline ..................................................................................................................... 7 2 Background ....................................................................................................................... 8 2.1 Gas sensing technology for wearable air pollutant monitoring device ............................. 8 2.1.1 Gas sensor requirements for wearable system implementation ................................... 8 2.1.2 Analysis of gas sensing technologies for wearable system.......................................... 9 2.1.3 Electrochemical sensing technology .......................................................................... 12 2.1.4 Summary of motivation for RTIL-based electrochemical gas sensor ........................ 15 2.2 Electrochemical interface IC ............................................................................................ 15 2.2.1 Potentiometric technique ............................................................................................ 16 2.2.2 Amperometric technique ............................................................................................ 16 2.2.3 EIS technique .............................................................................................................. 19 2.2.4 Summary of electrochemical IC circuit design .......................................................... 20 2.3 Integration approaches for gas sensor microsystem ........................................................ 21 2.3.1 Comparison between hybrid and monolithic approach.............................................. 21 2.3.2 Analysis of monolithic gas sensor implementation approach .................................... 22 2.3.3 Summary of integration approach for electrochemical gas sensor microsystem ...... 28 3 Development of RTIL-based Sensor for Microsystem Platform ............................... 30 3.1 Overview .......................................................................................................................... 30 3.2 Development of fast-response RTILs-based electrochemical sensor .............................. 30 3.2.1 Sensor’s structure design for fast response ................................................................ 30 3.2.2 Fabrication of PEoPM sensors ................................................................................... 33 3.2.3 Test of PEoEM sensors ............................................................................................... 36 3.3 Miniaturization approach of PEoPM sensor arrays ......................................................... 41 3.3.1 Fabrication of miniaturized PEoPM sensors .............................................................. 41 3.3.2 Package of miniaturized PEoPM sensor arrays ......................................................... 46 iv 3.3.3 Test of miniaturized PEoPM sensor arrays ................................................................ 48 3.3.4 Integration of PEoPM sensor array within a wearable monitoring system ............... 53 3.4 Development of flexible PEoPM sensors ........................................................................ 53 3.4.1 Fabrication of flexible PEoPM sensors ...................................................................... 54 3.4.2 Test of flexible PEoPM sensors .................................................................................. 58 3.5 Conclusion ........................................................................................................................ 63 4 Development of a Compact Low-Power Amperometric Instrumentation with Current-to-Digital Readout ........................................................................................... 64 4.1 Overview .......................................................................................................................... 64 4.2 Background ...................................................................................................................... 65 4.2.1 Electrochemical sensors and equivalent circuit models ............................................. 65 4.2.2 Traditional amperometric instrumentation ................................................................. 67 4.3 Compact amperometric instrumentation design .............................................................. 70 4.4 Performance analysis ....................................................................................................... 75 4.4.1 Performance difference from traditional instrumentation .......................................... 75 4.4.2 Second-order effects of the sensor equivalent circuit model ..................................... 77 4.5 Results .............................................................................................................................. 79 4.5.1 CCDAI implementation .............................................................................................. 79 4.5.2 Experimental results ................................................................................................... 81 4.5.3 Performance comparison ............................................................................................ 84 4.6 Conclusion ........................................................................................................................ 86 5 Development of a CMOS Monolithic Electrochemical Gas Sensor Microsystem ... 87 5.1 Overview .......................................................................................................................... 87 5.2 System architecture .......................................................................................................... 87 5.2.1 Electrochemical sensor system ................................................................................... 87 5.2.2 CMOS monolithic sensor microsystem concept ........................................................ 89 5.3 CMOS design ................................................................................................................... 91 5.3.1 Potentiostat ................................................................................................................. 91 5.3.2 Amperometric readout circuit..................................................................................... 92 5.4 Monolithic sensor fabrication .......................................................................................... 93 5.5 Results .............................................................................................................................. 96 5.5.1 Test setup .................................................................................................................... 96 5.5.2 Electrical experimental results ................................................................................... 97 5.5.3 Chemical experimental results ................................................................................... 98 5.6 Conclusion ...................................................................................................................... 105 6 Summary, Contributions and Future Work .............................................................. 106 Summary ........................................................................................................................ 106 6.1 v 6.2 6.3 Contributions .................................................................................................................. 107 Future work .................................................................................................................... 108 APPENDICES ............................................................................................................................. 111 Appendix A: EIS equivalent circuit modeling ......................................................................... 112 Appendix B: Fabrication procedure for PEoPM sensors......................................................... 118 BIBLIOGRAPHY ...................................................................................................................... 125 vi LIST OF TABLES Table 2.1. Summary of gas sensing technologies. ····················································· 12 Table 2.2. Summary of monolithic gas sensors. ······················································· 28 Table 3.1. Average sheet resistance and standard deviation with different width configurations. · ···································································································· 44 Table 4.1. Area occupation of IC modules in ON 0.5µm CMOS fabrication process and the comparison between the traditional amperometric instrumentation circuit and the CCDAI. ····· 85 Table 4.2. Power consumption of IC modules in ON 0.5µm CMOS fabrication process and the comparison between the traditional amperometric instrumentation circuit and the CCDAI. ····· 85 Table 5.1. Performance summary of the CMOS interface circuit. ·································· 99 Table 5.2. Off-CMOS and on-CMOS electrochemical RTIL sensor measured parameters and calculated volume and height. ············································································ 104 Table 5.3. Off-CMOS and on-CMOS electrochemical RTIL sensor noise level comparison. ······························································································· 104 vii LIST OF FIGURES Figure 1.1. Concept of the wearable platform for air pollutant monitoring and data collection. 3 Figure 1.2. Architecture of a portable air pollutant monitoring device. ····························· 5 Figure 2.1. Simplified block diagram of circuits for potentiometric measurements. ············· 16 Figure 2.2. Simplified block diagram of circuits for amperometric measurements. ·············· 17 Figure 2.3. The structure of basic potentiostat. ························································ 18 Figure 3.1. (a) Conventional sensor structure: response time is slow due to slow gas diffusion through RTIL; (b) RTIL-in-permeable-membrane structure: response time is improved by a controllable thin electrolyte layer; (c) Porous-electrodes-on-RTIL structure: response time is improved by bypassing gas diffusion in RTIL. (d) Electrodes-on-permeable-membrane structure: response time is improved due to fast gas diffusion in the permeable membrane. ················· 32 Figure 3.2. (a) Image of RTIL [C4mpy][NTf2] droplet on porous PTFE. The contact angle is 95˚. (b) Image of RTIL [C4mpy][NTf2] droplet on gold-coated porous PTFE. The contact angle is 18˚. ·································································································· 34 Figure 3.3. PEoPM sensor fabrication process. A stainless steel mask was mounted against the porous PTFE and then a gold film was deposited. After removing the mask, patterned planar gold electrodes were formed. Finally, RTIL was coated on the electrodes to form the sensor. ········· 35 Figure 3.4. Photograph of fabricated PEoPM device. The WE and CE are interdigitated with finger widths of and the gaps of 200 µm. The O-ring isolates the sensor area from the contact pads. ·································································································· 36 Figure 3.5. Cross-sectional view of the gas testing chamber. Separate inlets and outlets on the top and bottom of the chamber allow gas to be introduced along two separate paths to effectively characterize two different sensor structures, where path1 corresponds to a conventional sensor structure shown in Figure 3.1(a) and path 2 corresponds to the PEoPM sensor structure shown in Figure 3.1(d).·································································································· 37 Figure 3.6. Impedance amplitude |Z| spectrum from 0.1 Hz to 10 kHz recorded from an RTIL-based PEoPM sensor presenting under a SO2 environment. SO2 concentration varies from 0 viii to 25ppm. Obvious impedance amplitude change to different SO2 concentration was observed, with a stronger response below 100Hz. ··································································· 38 Figure 3.7. SO2 calibration curve for RTIL-based PEoPM sensor. Blue dots represent the recorded -Zim values at 10Hz. Red curve represents the fitting curve follows Equation 3.1. The 2 fitting curve achieves high least-squares correlation coefficients (R =0.9957). ···················· 39 Figure 3.8. Transient response of RTIL-PEoPM sensor to 25 ppm SO2, with gas flow from two different paths (see Figure 3.5.) representing a conventional sensor structure with 56 s full scale response and a new rapid-response structure that reduces response time to 20 s. ·················· 40 Figure 3.9. Profile scan by Dektak3 Surface Profiler. The photoresist AZ4620 covered the rough surface of porous PTFE and a good step was formed. ·················································· 43 Figure 3.10. Miniaturized RTIL-based PEoPM sensor fabrication process. ························ 45 Figure 3.11. Photograph of fabricated miniaturized PEoPM device. The WE and CE are interdigitated with finger gaps of 100µm. The O-ring isolates the RTIL area from the contact pads. ·································································································· 46 Figure 3.12. (a) Illustration of the parts assembly for PEoPM sensor array. First, four electrodes-patterned PEoPM sheets were placed on wire-routed PCB substrate, with pads connected to wire contacts on PCB by conductive epoxy. Then O-rings were attached on a windows-opened PCB lib and mounted on the wire-routed PCB substrate by tighten bolts, to create RTIL reservoirs. Finally, [C4mpy][NTf2] were dropped in the reservoirs to create RTIL sensing layer. (b) Photographs of assembled PEoPM sensor array from front, back and side views. The sensor array occupies 1.55” x 1.25” including the connector to be electrically connected with sensor interface circuit. ······················································································ 48 Figure 3.13. CH4, SO2, NO2, O2 calibration curves for RTIL sensors from EIS test data at 10Hz. ·································································································· 51 Figure 3.14. Selectivity experiment. 20.4% O2, 3% CH4, 10 ppm SO2, and 10 ppm NO2 were purged into the test chamber sequentially in four steps and the sensor array’s impedance responses ix were recorded in each step. Taking the impedance response to O2 as 100% signal S, the sensor array responses to all four gases are normalized as ΔS/S (%) and plotted. ·························· 52 Figure 3.15. 2.6” × 2” × 1” prototype wearable air pollutant monitoring system with miniaturized sensor array, multi-electrochemical instrumentation board (MEIB), and a µ-Controller (µC) board. ·································································································· 53 Figure 3.16. Coating experiments illustration. Case1 illustrates small volume pure RTIL coating (Vol.=0.2 μL) result. RTIL droplets were coated only electrodes surface. Case2 illustrates large volume pure RTIL coating (vol.=1 μL) result. RTIL droplets were coated on electrode surfaces as well as the gap between electrodes. Case3 illustrates ethanol/RTIL mixture coating (Rethanol/RTIL=2, RTIL’s vol.=0.2 μL) result. The initial mixture droplet was coated the electrodes area as well as the gap between electrodes. After ethanol removal, RTIL still occupied the initial droplet region. Case4 illustrates ethanol/RTIL mixture coating (Rethanol/RTIL=4, RTIL’s vol.=0.2 μL) result. The initial mixture droplet filled the pores inside the porous PTFE as well as covered electrodes surface. After ethanol removal, RTIL still stayed in the porous PTFE.·················· 55 Figure 3.17. (a) Photograph of microfabricated flexible PEoPM RTIL-based sensor. (b) Photograph of convex bend of the PEoPM RTIL-based sensor. (c) Photograph of concave bend of the PEoPM RTIL-based sensor. ············································································ 57 Figure 3.18. Gas test setup and position of the sensor within the flow cell.························· 58 Figure 3.19. Cyclic voltammograms (scan rate = 500 mV/s) for the reduction of O2 at WE. O2 concentration varies from 0-21%. O2 reduction current peak was observed at -1.4V vs Au. The relationship between peak current and O2 concentration from 0% to 21% is linear. ··············· 59 Figure 3.20. Current vs. time curve at various O2 concentrations when the potential is held at –1.4 V vs Au. N2 is the background gas. O2 concentration steps up from 0% to 21% and steps down from 21% to 0%. ····························································································· 61 Figure 3.21. O2 calibration curve for RTIL sensor using data extracted from Figure 3.19.······· 61 x Figure 3.22. Constant-potential current response measured over three cycles of alternate exposure to 1% O2 and pure N2 flow at an applied potential of -1.4 V. ········································· 62 Figure 4.1. Equivalent circuit model of electrochemical sensor cell. (a) Randles model (b) Complete model considering both AC and DC stimulus. (c) Simplified model for circuit analysis. ·································································································· 66 Figure 4.2. Schematic of a traditional amperometric instrumentation circuit including potentiostat and current-mode ADC. ······································································ 68 Figure 4.3. Waveforms of the current on the integrator input Iint, the voltage on the integrator output Vint, and the digital output of the comparator Dn. ·············································· 69 Figure 4.4. Schematic of the electrochemical sensor system. It consists of the traditional amperometric instrumentation circuit and the simplified electrochemical sensor equivalent circuit model. ·································································································· 71 Figure 4.5. Derivation of the instrumentation topology. The input current source is folded to parallel connect with integrator capacitor. ································································ 72 Figure 4.6. Schematic of the modified amperometric instrumentation circuit with sensor equivalent circuit model. ···················································································· 73 Figure 4.7. Schematic of the optimized amperometric instrumentation circuit with sensor equivalent circuit model. ···················································································· 74 Figure 4.8. Schematic of CCDAI with the simplified sensor equivalent circuit model. ········· 74 Figure 4.9. Vint waveform illustration when considering Rs in the equivalent circuit model. ·· 77 Figure 4.10. Illustration of Isens in time domain. ························································ 79 Figure 4.11. Test setup for electrical and chemical experiment. ······································ 80 Figure 4.12. A hysteresis comparator realization with adjustable upper/lower bound. ············ 80 Figure 4.13. DNL and INL of the CCDAI. Both DNL and INL in the current range are better than -49dB, implying an 8 bit of effective resolution. ························································ 81 xi Figure 4.14. The Faradaic current generated by 6 mM of potassium ferricyanide as function of time when VWE-RE=190mV. Red line represents data recorded by CHI760C and blue line represents data recorded by CCDAI. ······································································ 83 Figure 4.15. The Faradaic current recorded by the CCDAI at VWE-RE=190mV as function of time for 0- 6 mM of potassium ferricyanide. ··································································· 83 Figure 4.16. Calibration curve of Faradaic current vs potassium ferricyanide concentration. The current values were the average values from 200s to 300s. Fitting curve was presented as a straight 2 line. R values of the fitting line are 0.991 and 0.996 for the data acquired by CCDAI and CHI760C, respectively. ······················································································ 84 Figure 5.1. Block diagram of the electrochemical sensor system. ·································· 88 Figure 5.2. Conceptual illustration of the CMOS monolithic electrochemical gas sensor microsystem. ·································································································· 91 Figure 5.3. Schematic of the monolithic microsystem, consisting of the amperometric instrumentation circuit and the RTIL sensor. The green rectangle block illustrates the clock waveform. ·································································································· 92 Figure 5.4. Post-CMOS pocess flow for on-chip RTIL sensor fabrication: Shipley 1813 photoresist is spin-coated (a) and developed (b). Titanium and gold in deposited by PVD (c). Photoresist is then rinsed off to leave the electrode (d). CMOS chip with on-chip electrode is wire bonded to package (e). A droplet of RTIL is casted on the electrodes (f). ··························· 95 Figure 5.5. Die photographs of CMOS chip fabricated by foundry and on-CMOS RTIL sensor formed by post-CMOS fabrication. ········································································ 95 Figure 5.6. Chemical test setup. ·········································································· 97 Figure 5.7. The rail-to-rail opamp output voltage range and linearity error as a function of input voltage. The absolute error of the output voltage is less than 0.05% over the range from 0.03 V to 4.90 V. ·································································································· 99 Figure 5.8. Amperometric readout output range and linearity errors as a function of input current when fs = 100 kHz, Vref2 = 1.3 V. In the range of -1 µA to 1 µA, the absolute linearity errors are less than 4 nA. ································································································ 99 Figure 5.9. Output of amperometric readout circuit for on-CMOS RTIL sensor response to oxygen. Oxygen concentration varies from 0 - 21% with a step of 2.1%. Oxygen exposures were xii 10 s in duration, preceded and followed by exposure to pure nitrogen for 30 s. The amperometric circuit was set to Vref1=2.1 V and Vref2=1.3 V (the sensor was applied at -0.8V respectively), and fs=100 kHz. ································································································ 101 Figure 5.10. Oxygen calibration curve for monolithic RTIL gas sensor using data extracted from 2 Figure 5.8. A high level of linearity (R = 0.995) was achieved. ···································· 101 Figure 5.11. Repeatability test. Constant-potential current response measured over five cycles of alternate exposure to 4.2% oxygen and pure nitrogen flow at an applied bias of -0.8 V. The standard deviation of the five voltage data is 89 µV and corresponding repeatable LOD of oxygen is 0.06%. ································································································ 102 Figure 5.12. Off-CMOS and on-CMOS sensor response time (10% - 90%) characterized by amperometric readout circuit. ············································································ 104 Figure A.1. Randle equivalent circuit model. ························································· 112 Figure A.2. The equivalent circuit model when bias is much lower than over-potential. ······ 115 xiii 1 Introduction 1.1 Motivation 1.1.1 Air pollutants effect on human health and research method Airborne pollutants are known to threaten human health and safety, causing discomfort, illness, and even death, particularly among susceptible individuals such as those with pre-existing cardiovascular or respiratory problems as well as infants and the elderly [1, 2]. Exposure to air pollutants consistently ranks among the leading causes of illness and mortality. US Environmental Protection Agency has released National Ambient Air Quality Standards (NAAQS) [3] for the evaluation of the air pollution effects to public human health, where air quality is quantitatively identified by six principle pollutants including NOx, CO, O3 and SO2 from short-term and long-term aspects. The most pervasive methodology used to study the effects of air pollution on human health is epidemiology [1-3]. Epidemiological studies model the impact of air pollutants by illuminating statistical associations between levels of air pollutants and impacts on public human health both in short and long term. Such studies heavily rely on air pollutant data (species and concentration), which are acquired from stationary air monitoring sites. 1 1.1.2 Air pollutant monitoring and data acquisition platform Although epidemiology reveals statistical correlation between air pollutant exposure and human health, current research cannot produce precise models on the impacts of exposure to air pollutants on personal health. There are several limitations of existing data acquisition platforms that hinder the research [1]. Firstly, because of the high cost and bulky size of multi-gas analyzers using for air pollutant data recording, only a limited number of stationary sites are distributed in certain regions. Given the density of the distributed stations, the spatial resolution needed to track variations of pollutant concentrations within the region of interest is not usually provided. Secondly, a limited number of stationary sites are not able to provide pollutant data covering human-scale microenvironments. In addition, different daily activity patterns of varied sub-group of populations—for example adults and children—lead to different levels of exposure. Fixed monitoring sites fail to capture such differences. Given all these limitations, ambient levels of air pollutants were shown to be poor predictors of personal exposure, in particular, of acute exposure. The lack of an effective platform to acquire sufficient air pollutant data, in terms of both temporal and spatial resolution, remains the key obstacle to fully understand the health effects of pollutant. The size, cost and immobility of the platform limits its distribution, thus is the bottleneck of the research on health effects of air pollutants. To overcome the limitations of existing data acquisition platforms, it is desirable to have an alternative platform, that can be worn by many people, collect data continuously and automatically 2 Figure 1.1. Concept of the wearable platform for air pollutant monitoring and data collection. (For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this dissertation.) upload data to data clouds, thus providing data at high resolution both temporally and spatially as shown in Figure 1.1. In addition to providing information for epidemiology, this platform would also provide preventative alerts and managed care services to individuals susceptible to the effects of exposure. 1.2 Challenge To realize a wearable platform for air pollutant monitoring, individual monitoring device in the platform should have features as follows. To be able to monitor air pollutants, the device should be sensitive to the air pollutants in the range of interest, be reliable, and respond in real time. To provide preventative alerts and managed care services, the device should have signal processing and data storage. To upload air pollutant data to data clouds, the device should have 3 network function, like WiFi, Bluetooth or GPRS. To be able to be distributed to many people, the device should be inexpensive. To be easy to be worn daily, the device should be small and light like a cellphone or even a wrist watch. To be powered up to days or even months by battery, the device should be highly power efficient. Recently, a few portable air pollutant monitoring device prototypes [4-8] have been developed to realize the wearable platform concept. Their typical architecture is shown in Figure 1.2. It consists of four parts: (1) Multiple-gas sensor array transduces air pollutant information (species/concentration) to electrical signal; (2) Sensor interface circuit drives the sensor array as well as readouts the electrical signal; (3) µController/ARM module digitizes and processes the sensor signal; and (4) Bluetooth/WiFi/GPRS module uploads the data to the data clouds. Those device prototypes meet some criteria of the wearable platform. They are capable of monitoring single or multiple air pollutants and uploading data to the Internet. However, they are impractical to distribute to many people and continuously collect data due to three common drawbacks: high product cost, high power consumption, and relative large size. Each device costs several hundred dollars. Their power consumption is around 100mW to 1W, limiting their battery-powered lifetime only to a few hours. Although the devices are portable, they are too large to be wearable, like the size of a smart phone or a wrist watch. The power/size/cost drawbacks of existing gas monitors can be analyzed in terms of their general architecture. Thanks to modern digital and RF IC technologies accelerated by smart phones/tablets, the µController/ARM module and Bluetooth/WiFi/GPRS module are typically 4 Figure 1.2. Architecture of a portable air pollutant monitoring device. implemented using commercial IC chips with low size, low cost and low power. In contrast, currently commercial sensor arrays and sensor interface circuits are of high cost and large size. Therefore, due to the underdeveloped sensor array and sensor interface circuit technologies, existing portable air pollutant monitoring device prototypes cannot directly be used for the wearable platform. Mirroring the success of embedded camera sensor units in smart phones, if the price of the sensor unit (including sensor interface circuit) falls below $1 and the size smaller than 3 1cm , it is possible to develop a wearable microsystem with similar size to Google Glass or smart watches by incorporating sensors, sensor interface circuits and digital/RF modules together for air pollutant monitoring. To overcome the bottlenecks of the wearable platform, research efforts are needed that focus on the miniaturization of the sensor array and power-efficient interface circuits utilizing cost-effective fabrication processes. From the system point of view, several technical challenges remain in order to achieve a cost-effective gas sensor microsystem for long-term use. The first critical challenge lies in developing a miniaturized high-performance sensor device, because available commercial sensors cannot satisfy the wearable system’s requirement in terms 5 of power and size. The miniaturized sensor device should be able to sense the air pollutants in the range of interest in real time, with long lifetime, small feature, and low power. The second challenge lies in developing sensor interface electronic circuits. Traditionally, high-performance, high-cost, bulky instrumentation is used to characterize gas sensors in sensing technology development. When applied to field sensing, a low-power, low-cost IC solution is required. The third critical challenge lies in the integration approach for sensor devices and electronics circuits. The monolithic microsystem approach, that implements both sensor and electronics circuits in a single CMOS chip, is considered to be a cost-effective and size-compact solution. However it rouses technique challenges like process compatibility during implementation. 1.3 Goal The goal of this dissertation is to experimentally develop a wearable gas sensor microsystem utilizing a sensing technology and integration approach that satisfies all the requirements, including performance, cost, and size. Scientific contributions are expected in the areas of miniaturization of high performance gas sensor arrays, compact sensor interface circuits, and sensory microsystem integration. 6 1.4 Thesis outline Chapter 2 describes the background of this research. Options of potential gas sensing technologies suitable for wearable gas sensor microsystem are discussed. With the best sensing technology selected, options of potential electronics are discussed. The possible integration approaches are discussed with a review of current research progress. Chapter 3 reports the sensor device development utilizing the selected sensing technique. Its miniaturization approach is also presented. Chapter 4 introduces an optimized compact sensor interface circuit implementation tailored to the chosen sensing methodology. Chapter 5 introduces the development of a monolithic sensor microsystem platform that enables highly-miniaturized wearable air pollutant monitoring device. Conclusions and contributions are given in Chapter 6. 7 2 Background 2.1 Gas sensing technology for wearable air pollutant monitoring device 2.1.1 Gas sensor requirements for wearable system implementation To provide information for epidemiology to build accurate personal exposure model, and provide preventative alerts and managed care services to individuals susceptible to the effects of exposure, a wearable air pollutant monitoring platform is demanded. To realize a wearable platform for air pollutant monitoring, individual monitoring devices should meet following criteria. (1) Sensitive to the air pollutants (NO2, SO2, O3, CO, volatile organic compounds (VOC), etc.) in the range of interests, be reliable and respond in real time. (2) Signal processing and data storage capability. (3) Networking capability. (4) Cost effective to be widely distributed. (5) Miniaturized size and light weight equivalent to a cellphone or even a wrist watch. (6) Power efficient that are can be powered up to days or even months by battery. Power range from mW to µW is desirable. To satisfy the device criteria, the gas sensor in the device should meet following requirements. 8 (1) Sensitive to the air pollutants (NO2, SO2, O3, CO, VOCs, etc) in the range of interest with good selectivity. (2) Good reliability. It should function for months to years without extract maintenance. (3) Rapid response time to achieve high temporal resolution, sampling rate up to 10Hz. 3 (4) Miniaturized size to be fit into cellphone or even wrist watch. Size around cm or even 3 mm is desirable. (5) Low power consumption around µW is desirable. 2.1.2 Analysis of gas sensing technologies for wearable system To develop a miniaturized gas sensor that meets all the requirement, prevalent gas sensing technologies should be analyzed, and the most promising sensing technology would be chosen for the sensor development. The five types of prevalent gas sensing technologies are: gas chromatography (GC)-based [9], non-dispersive infrared (NDIR) [10], metal oxide semiconductor [11], conductive polymer [12], and electrochemical sensing technology [13]. GC-based sensing technology [9] refers to the gas sensing technology that utilizes a semi-selective gas sensor and adopts GC to enhance selectivity. GC plus mass spectrometry (MS) [14] and flame ionization detector (FID) [15] are standard gas analyzers for stationary air pollutant monitoring sites [16]. GC-based gas analyzers achieve high sensitivity, selectivity, and reliability. However, traditional GC-based gas analyzers are bulky and expensive, require complex operations and regular maintenance. Although development of micro GC [17, 18] , micro MS [19, 20] and 9 micro FID [21] have been reported during last decade, the GC’s high degree of complexity and sophistication still results high production cost and maintenance effects. In addition, the operation of GC-based gas analyzers involves a gas sampling process and a pre-concentration process, which brings significant drawbacks. The gas sampling process is time consuming, usually taking minutes to output one dataset. Thus the sampling rate is pretty low. The gas pre-concentration process needs a high-power-consumption heater to release absorbed gas analytes. Thus the power consumption of GC-based gas analyzers is inefficient. NDIR sensing technology [10] detects gas by determining the absorption of an emitted infrared light source through a certain air sample. Commercial NDIR sensors can detect air pollutants – CO, NO, SO2, O3 and VOCs [22]. NDIR sensors exhibit good sensitivity and good stability. However, they necessitate a relatively complicated optical system, making it expensive, bulky and high power consumption (hundred mW). Metal oxide semiconductor gas sensing technology [11] has been developed since 1962 and demonstrated for a variety of air pollutants—CO, NO, NO2, O3 and VOCs [22]. The sensing principle is: the film resistance changes when gases are absorbed or combust with catalyst material. By choosing different metal oxide semiconductor materials and doping/coating different catalyst materials, metal oxide semiconductor sensors exhibit good selectivity to gas analytes. Other advantages include light weight, miniaturized size, quick response time, and low cost. However, metal oxide semiconductor gas sensor requires operating at very high temperature (usually >250°C), and thus consume a lot of power. Even a miniaturized sensor could consume 10 power around 200 mW [23]. In addition, although they have long operation lifetime and reasonable parameter stability, they degrade as they are used. Conductive polymer sensing technology [12] uses different types of conductive polymer to absorb or react with a specific gas, causing a change of the polymer film resistance. It has been reported as capable of sensing CO, NO2 and a few VOCs in the air pollutant category. Conductive polymer based sensors consume low power given their room-temperature operations. They exhibit high sensitivity and short response time to analytes. They have low fabrication cost because of a simple fabrication process and are easy to miniaturize. However, they suffer low selectivity, long-term instability and irreversibility. Electrochemical sensing technology [13] utilizes electrochemical reaction to generate an electrical signal (current or potential) to sense gases. Electrochemical sensors have been proven to be effective for detecting environmental gases such as CH4, CO, CO2, NO, NO2, SO2, H2, O3, O2 and VOCs [22, 24-27]. They exhibit good selectivity, low power consumption (in liquid electrolyte), wide dynamic range and low production cost. Their response time and reliability vary with the electrolyte. They have the potential to be miniaturized with the newly-developed electrolytes. In all, the five gas sensing technologies are qualitatively summarized in Table 2.1 in terms of gas sensor criteria for microsystem implementation. Although GC-based sensing technology is capable of sensing all air pollutants of interest, and exhibits excellent selectivity and reliability, its high cost, slow response, and high power consumption exclude itself from being an ideal candidate 11 for wearable air pollutant monitoring microsystem. Apart from GC-based sensors, only electrochemical sensors are sensitive to all the air pollutants of interest and they exhibit the best comprehensive strength to satisfy all requirements for air pollutant monitoring microsystem. Therefore, the electrochemical sensing technique is chosen to develop miniaturized sensor arrays for gas sensor microsystems. 2.1.3 Electrochemical sensing technology To develop an electrochemical gas sensor that satisfies all the requirements for the air pollutant monitoring microsystem, electrochemical sensing technology is reviewed in detail. Technique challenges of the wearable electrochemical sensor development are thereby revealed. Electrochemical gas sensing [13] is based on the measurement of an electrical signal caused by chemical reaction in which the gas analyte is involved. The reaction takes place when the gas diffuses to the interface of an electrode and a supporting electrolyte and a certain electrical Table 2.1. Summary of gas sensing technologies. Sensing technology Detectable air pollutants GC/MS or GC/FID Metal oxide Conductive Electrochemical semiconductor polymer NDIR CO, NO, NO2, CO, NO, SO2, O3, and VOCs Selectivity Excellent Reliability Excellent Low cost Bad Fast response Bad Miniaturized size Good Power consumption Bad CO, NO, CO, NO, NO2, CO, NO2, SO2, O3 NO2, O3 and and VOCs SO2, O3, and and VOCs VOCs VOCs Good Good Bad Good Good Moderate Bad Moderate Bad Good Good Good Good Good Good Moderate Bad Good Good Good Bad Bad Good Good 12 condition is satisfied. This type of sensor has been utilized in industry over the last 50 years. Electrochemical gas sensors provide one of the lowest-power approaches combined with outstanding performance that includes high sensitivity and good selectivity. They are sensitive to a wide range of air pollutants such as CH4, CO, CO2, NO, NO2, SO2, H2, O3 and O2. Furthermore, the instrumentation needed for such sensors can readily be implemented as a single microelectronics chip suitable for wearable/portable devices [28-30]. Different planar electrodes configuration are adopted for different working modes. Concentric electrodes are used for amperometric mode to achieve a uniform electrical field. Interdigitated electrodes are used for electrochemical impedance spectroscopy (EIS) mode to minimize the solution resistance. In terms of electrolyte type, there are two types of electrochemical gas sensors: solid-electrolyte sensors and liquid-electrolyte sensors. Solid-electrolyte electrochemical sensors are utilized in the automotive industry to detect exhaust gases. Because they can only function at high temperature, to achieve sufficient ion mobility to sense gases, heaters are required if the solid-electrolyte electrochemical sensors are applied to air pollutant monitoring. This requirement brings extra power consumption and is not favorable. In contrast, liquid-electrolyte electrochemical sensors are one of the lowest-power gas sensors because of room temperature operation. They can detect a variety of gases (e.g. O2, CO, SO2, H2S, NO2) by employing conventional solvents (e.g. H2SO4/H2O mixtures, or organic solvents such as acetonitrile or propylene carbonate). However, traditional liquid-electrolyte electrochemical sensors cannot 13 survive in drastic temperature changes and cannot function in extremely dry or humid conditions; such conditions exist in normal circumstances. In addition, the lifetime of such liquid-state electrolyte sensors is often determined by the volume of the electrolyte in the sensor and how quickly the electrolyte dries up. Such features make traditional liquid-electrolyte electrochemical sensors unfavorable for miniaturization. To overcome the drawbacks of the traditional liquid-electrolyte electrochemical sensors, room temperature ionic liquids (RTILs) have been adopted as a new type of liquid electrolyte recently [31-33]. RTILs are nonvolatile and conductive compounds consisting entirely of ions [31]. Given the special properties, RTILs are a promising electrolyte for robust electrochemical gas sensors that can operate in extreme conditions. On the one hand, RTIL-based electrochemical gas sensors inherit the advantages of liquid-electrolyte electrochemical gas sensor. RTILs can function at room temperature. Thus, power demands are greatly reduced by eliminating the heater required by many other sensors. On the other hand, RTILs overcome the drawbacks of traditional liquid-electrolyte electrochemical gas sensor. RTIL has negligible vapor pressure and high thermal stability. It features longer lifetime, structure that does not require a permeable membrane to seal liquid, and capability to function in a very small volume. The selectivity of an electrochemical gas sensor is introduced by the electrochemical potential applied to the sensor. Therefore, an electrolyte that can be sustained at a wide range of electrochemical potential has a possibility to sense many gases. Thanks to their wide electrochemical potential windows, RTILs are capable of sensing a variety of gases, including oxygen [34-39], ambient toxic gases (e.g. NO2 [40], NO [41], 14 NH3 [42], H2S [43]) and VOCs [44, 45]. However, as the viscosity of RTILs is typically larger than traditional electrolytes by 1-2 orders of magnitude, the diffusion coefficients of gaseous analytes are significantly lower in RTILs. Because the reaction only takes place when the gas reaches the interface of the electrode and electrolyte by diffusion, the response of RTIL-based sensor is slow in a traditional electrochemical cell structure where a thick layer of electrolyte is usually deployed between the electrodes and atmosphere. 2.1.4 Summary of motivation for RTIL-based electrochemical gas sensor In all, the electrochemical sensing technique demonstrates overall good performance, and thus has the potential to meet all requirements for air pollutant gas sensor microsystems. Choosing RTIL as electrolyte, the electrochemical sensors overcome drawbacks of traditional liquid-electrolyte electrochemical sensors. The challenge of slow response requires an innovation in sensor structure. In addition, an approach to sensor array miniaturization is demanded. 2.2 Electrochemical interface IC Because an RTIL-based electrochemical gas sensor array has been selected for microsystem implementation, a corresponding sensor interface circuit should be developed to meet system criteria as well. Traditionally, high-performance, high-cost, bulky instrumentation is used to characterize gas sensors in sensing technology development. When applied to wearable microsystems, a low-power, low-cost IC solution is required. Although no IC solution for 15 electrochemical gas sensors has been reported, ICs for other electrochemical applications [28-30] can be referenced for developing a gas sensor interface IC, given the same working principle. Based on implementation methods, electrochemical instrumentation circuits can be classified into three types: potentiometric, amperometric, and EIS technique. In this section, a brief review on existing electrochemical instrumentation circuits implemented in CMOS IC is presented. 2.2.1 Potentiometric technique The potentiometric technique measures potential between a working electrode (WE) and a reference electrode (RE) when a controlled current is passed between the WE and a counter electrode (CE) [46] (Figure 2.1). Due to relatively large error caused by double-layer charging effects, this method is seldom used for sensors. No IC instrumentation circuit implementing potentiometric sensing has been reported. 2.2.2 Amperometric technique The amperometric technique measures current between a CE to a WE with a current readout circuit when a controlled potential (sustained by potentiostat) is applied between RE and WE as Figure 2.1. Simplified block diagram of circuits for potentiometric measurements. 16 shown in Figure 2.2 [46]. In the amperometric sensor, the current is associated with an oxidation and/or reduction reaction involving the analyte. The controlled potential typically is a constant or ramp potential. Applying constant or ramp potential stimulus respectively, chronoamperometry and cyclic voltammetry technologies are widely used in enzymes, DNA and protein detections. In order to meet different demands for sensor applications, numerous CMOS integrated potentiostats and current readout circuits have been developed with various functions and performances in the past decades. A basic potentiostat is implemented by an amplifier as shown in Figure 2.3. By incorporating amplifier-based adder, multiple stimulus signals can be added to the electrochemical sensors. By adjusting stimulus to the WE side, a hardware-efficient bi-potentiostat scheme was implemented in CMOS IC for redox-enzyme-based biosensors [47]. To improve the output swing and dynamic range, a fully-differential potentiostat was developed to be capable of being fabricated in a low-voltage CMOS process [48]. A current-mirror-based potentiostat was presented in [49] to achieve lower power consumption and occupy smaller area. Figure 2.2. Simplified block diagram of circuits for amperometric measurements. 17 Figure 2.3. The structure of basic potentiostat. A basic current readout circuit is composed of a current-voltage convertor and a voltage-mode analog-to-digital converter (ADC). The former part typically is implemented by a resistive feedback amplifier or an integrator; the later part typically is implemented by a dual-slope ADC [50, 51]. An advanced readout circuit approach adapts current-mode ADC to directly digitize the current [52]. In the current readout circuits design, the key point is to improve the detection resolution. The main barriers are 1/f noise and thermal noise in the circuits. 1/f noise is the dominant noise source for DC (or low frequency for the low scan rate case in CV) current detection circuits. Several techniques, such as chopping and correlated double sampling (CDS) can eliminate 1/f noise. In [53], the entire current readout chain utilized CDS to reduce 1/f noise as well as amplifier offset. To reduce thermal noise, one solution is to use a longer integration time [54, 55], in which a switched-capacitor first-order single-bit Σ-Δ modulator was implemented in the current-to-digital converter. However, the single Σ-Δ technique reduces the thermal noise by 18 sacrificing the speed of the converter. In [52], a semi-synchronous Σ-Δ analog-to-digital conversion algorithm, combines continuous time Σ-Δ with time-encoding machines (TEM) to improve the readout current resolution. For high throughput array microsystems, power consumption and area are important circuit design parameters to consider besides noise performance. A half-amplifier structure was developed to measure small currents without increasing the power consumption and layout area by eliminating amplifying circuitry [56]. 2.2.3 EIS technique In the EIS technique, a small-amplitude AC stimulus with known frequency content is applied to a sensor interface via a potentiostat. Current is then recorded for signal conditioning to extract a phase or amplitude shift, algebraically related to sensor interface’s impedance (real and imaginary components). When applying an analyte to the sensor, the impedance of the sensor changes respectively. This technique has been implemented in CMOS chips for applications such as protein sensors [30] and DNA sensors [57]. Regarding the extraction method, fast Fourier transform (FFT) algorithm and frequency response analyzer (FRA) based methods are typically used [58]. The FFT algorithm utilizes a broadband stimulus (e.g. a pulse) and computes the results at all frequency points simultaneously, producing an impedance spectrum. The composite signal source and the computationally intensive FFT require a digital signal processor (DSP) with extensive computational resources [59]. 19 Although a few studies have been conducted to implement FFT with analog circuits [60], it is impractical for microsystem implementations due to poor multiplier linearity and accuracy, and demands for preprocessing/pre-storing of the input data set. Alternatively, in FRA, a small amplitude (10mV) sinusoid voltage stimulus is applied between the two biosensor electrodes, and the current response is measured as the frequency of the stimulus voltage is swept. The FRA method can be realized with compact analog circuits [28-30]. Recently many FRA-based instrumentation circuits have been developed. The lock-in technique, which is a two phase reference coherent demodulation method, has been employed to extract the value of the magnitude and phase of impedance [28]. A circuit combined lock-in multiplier with Σ-Δ modulation reports ultra-low power consumption and high resolution [61]. A dual-slope ADC with lock-in function as well as an on-chip sinusoid signal generator was reported in [62]. To address the impediments of solution resistance, a differential FRA approach was applied in [63]. 2.2.4 Summary of electrochemical IC circuit design The literature review reveals that electrochemical IC circuit design is tailored to these specific applications. To maximize the sensor’s performance and meet system requirements, various kinds of measurement techniques, stimulus schemes, and readout topologies have been developed respectively. Therefore, customized electrochemical interface IC circuits should be optimized according to the gas sensor microsystem’s compact, low power, and low cost requirement. 20 2.3 Integration approaches for gas sensor microsystem To achieve highly miniaturized gas sensor microsystem, the gas sensor and sensor interface circuit should be integrated in a cost-effective and highly-compact way. Therefore, it is necessary to analyze existing integration approaches and select a path for the gas sensor microsystem integration, tailoring to electrochemical sensing technology. 2.3.1 Comparison between hybrid and monolithic approach Reviewing the past 20 years of academic and industrial efforts toward gas detector miniaturization, two main development pathways have emerged for the integration of gas sensor elements and sensor interface circuits. They are referred to as hybrid integration and monolithic integration [64]. In the case of hybrid integration, the sensor element and readout circuitry are formed by two separate elements that are assembled into a single module. Typically this involves packaged components combined on a printed circuit board (PCB). This approach permits sensor and circuitry to be fabricated using specific processes, materials, and thermal budgets that are tailored to the individual needs of each element. However, this approach has larger parasitic capacitance/resistance due to interconnection between two elements. It also has the problem of higher packaging cost and larger size than monolithic integration. In the case of monolithic integration, the sensor element and readout circuitry are formed within a single chip. Compared to hybrid integration, monolithic integration is superior in the 21 following perspectives. It is of smaller size. It has lower noise by eliminating interference coupled through wire bonds. It is also of higher sensitivity by minimizing parasitic capacitance/resistance, and of lower cost by reducing chip packaging. In conclusion, following the trend of miniaturization in electronic products, monolithic integration is a promising approach for wearable gas sensor microsystems. 2.3.2 Analysis of monolithic gas sensor implementation approach The monolithic integration approach has been applied to several sensor technologies including micro cantilever sensors, thermoelectric sensors, semiconductor sensors, optical sensors, chemiresistor sensors and chemical field effect transistor (ChemFET) sensors. Although no monolithic electrochemical gas sensor has been reported, it is valuable to understand all monolithic gas sensors’ working principles, fabrication process, pros, cons, etc. 2.3.2.1 Micro cantilever sensors Micro cantilever sensors are a type of gravimetric sensors which respond to the mass accumulated in a sensing layer. In the case of gas sensing, the sensing layer must be capable of selectively absorbing a target gas molecule. Micro cantilevers are fabricated on silicon substrate by using micro/nano-electromechanical systems (MEMS/NEMS) technology and then coated by a selective gas sensing polymer. The total mass of the cantilever changes when the sensing polymer absorbs gas molecules, causing alteration of resonant frequency [65, 66]. 22 The cantilever resonance is actuated by a heater through thermal-mechanical transduction, and is sensed by piezoresistor through mechanical-electrical transduction. A feedback-loop circuit controls the hotplate’s periodically fluctuating temperature to drive the cantilever vibration at its resonant frequency. Bulk micromachining is adopted to implement cantilevers by using wet etchants to isotropic etch the silicon substrate, leaving a p-doped (p+) region as cantilever. Coated with a specific gas-sensitive polymer, the micro cantilever is capable of detecting VOCs such as toluene, octane and ethanol. However, this type of cantilever gas sensor has two drawbacks. First, it has limited detection resolution. Second, power consumption is high. Due to the hotplate, power consumption can reach up to 10mW. 2.3.2.2 Thermoelectric sensors Thermoelectric sensors convert the temperature difference on junctions of two conductive materials to a voltage difference. After coating a gas sensitive polymer with conductive materials, the enthalpy change produced by ambient gas absorption or desorption can induce temperature change, thus create voltage change. In this way, the gas concentration is transduced to an electrical signal. Such sensors could be implemented by the CMOS process, utilizing polysilicon and aluminum as thermopiles. To fully implement these sensors, several extra components should be included. First, a heater is needed to generate a temperature difference, which can be easily realized by CMOS process. Second, a thermally isolated structure—such as a membrane and a 23 bridge—is required to enhance the temperature gradient between hot and cold junctions in a limited area and protect the circuit, which operates at room temperature. This requires an extra post-CMOS back-side etch. Third, a gas sensitive polymer needs to be coated on the thermopile to enable gas sensing. A monolithic thermoelectric sensor integrated with a pre-amplifier was reported to detect different VOCs [67]. An improved sensor structure with an N-well island membrane and a more accurate calibration was presented in a follow up paper [68]. However, this improvement does not solve the problem of extra power consumption introduced by the heater. Similar to micro cantilevers, extra power consumption remains one of the major drawbacks of thermoelectric sensors. 2.3.2.3 Semiconductor sensors Semiconductor sensors utilize metal oxide and noble metal additives as catalytic materials to detect combustible gases [69]. The combustible gases can be oxidized below the ignition temperature when catalytic materials are presented. The enthalpy change during the reaction induces a change in temperature and eventually in the resistance of the catalytic materials. Similar to the thermoelectric sensors, the implementation of monolithic semiconductor sensors also requires a heater, a thermally isolated island, sensing material coating, and readout circuit, either by CMOS or post CMOS process fabrication. Furthermore, to achieve a controllable high temperature (usually >400 °C), a hotplate with homogeneous temperature distribution, a temperature sensor and temperature control circuits are required. Such a monolithic sensor was 24 first reported in 2004 [2]. Later, an improved monolithic metal oxide catalytic sensor with post-CMOS fabricated temperature sensor that can operate at 500 °C was reported [23]. This compact array demonstrated improved analyte discrimination with a detection limit of 1 ppm CO and 100 ppm CH4. However, those monolithic semiconductor sensors suffer similar disadvantages as thermoelectric sensors, given the similarly high working temperature and complicated post CMOS process. 2.3.2.4 Optical sensors Compared to other sensing methods, optical techniques provide great selectivity by reading different properties of the electromagnetic waves such as frequency, phase, amplitude, and state of polarization. Photodiodes as optical sensors are readily fabricated by modern CMOS technique. However, since optical stimulus is difficult to integrate into monolithic chips, there is very little research on monolithic optical gas sensors. A technique utilizing bioluminescence was developed to implement a monolithic gas sensor called bioluminescent bioreporter integrated circuits (BBIC) [70-72]. This technique employs engineered bioluminescent bacteria, which can metabolize target compounds such as toluene and convert the energy directly into light. A conceptual BBIC microsystem is composed of porous light-tight closure, bioluminescent bacteria in a microenvironment, and IC with photodiodes on a chip. The IC chip detects the intensity of the luminescent light signal proportional to the target concentration. 10 ppb detection limit of toluene was reported. However, the disadvantages of the BBIC limit its application. First, 25 the target analyte is limited by the type of bacteria. Second, the BBIC is not applicable to real-time monitoring due to long response time ranging from minutes to hours. Third, the sensor needs a cultivated microenvironment, and the bacteria amount varies over time, causing signal drift. 2.3.2.5 Chemiresistor sensors Chemiresistor sensors rely on changes in the conductivity of a polymeric or organic sensitive film upon interaction with a gas analyte (absorption and desorption). Their operation temperature ranges from room temperature to 60 °C. Depending on the target analytes, the sensing materials vary from conductive polymers [73], carbon black polymers [74-76], thiolate-monolayer-protected gold nanoparticles (MPN) [77] to single wall carbon nanotubes (SWCNT) [78]. Because most sensing materials work at room temperature, island structure formed by back-side etch are not necessary and only on-chip surface modifications are involved. Generally, those materials are dissolved / dispensed into volatile organic solutions and sprayed onto the sensing area of the chips. On-chip electrodes were fabricated either by CMOS process or by post-CMOS metal electroplating, sputtering and thermal deposition. Special post-CMOS treatments includes annealing [74], dielectrophoretic assembly, parylene-C encapsulation [78]. 26 2.3.2.6 ChemFET sensors ChemFET sensors function similarly to MOSFET. The FET’s source-drain current changes with channel work function change, which is directly related to ambient gas concentration. Three types of ChemFET have been developed with CMOS-compactable process. The first type is called adsorption-porous-silicon-FET(APSFET) where the transistor channel is directly modified to a gas sensitive layer [79, 80].The second type is called Lundstrom-FET where the sensitive layer is deposited on top of the transistor channel [81]. The third type is called suspended-gate-FET (SGFET) where the gas sensitive layer is suspended at a certain distance above the channel [81]. All ChemFET fabrication processes include standard CMOS and post-CMOS processes. Regarding post-CMOS process, APSFET requires a BHF etching to create a porous silicon region for sensing; Lundstrom-FET requires palladium gate and SGFET need extra electrode deposition and suspended gate attaching. The gases sensed by ChemFET in air monitoring applications include NO2 and H2. The gas variety is mainly constrained by the sensing material. To date, limited metal (palladium in Lundstrom-FET and Pt in SGFET) could be deposited on silicon substrate and limited materials (boron in APSFET) could be doped in silicon substrate. 2.3.2.7 Summary of monolithic gas sensor technologies The monolithic gas sensors discussed above are summarized in Table 2.2. The summary is organized in terms of sensor technology, sensing material, detectable gases, and disadvantages. 27 Regarding performance, each type of sensor can only detect limited types of gases confined by the compactable sensing material it adopts. Besides the popular sensing materials like metal oxide and conductive polymer, others have limited potential to extend the monolithic platform to detect multiple gases. In addition, all sensors inherit disadvantages from the sensing materials they adopted, such as slow response to gases, sensor baseline drift, and extra power needed to heat up the sensor to the desired operation temperate. Lastly, our literature review leads us to conclude that no monolithic integration of electrochemical gas sensors has been reported. Table 2.2. Summary of monolithic gas sensors Sensing technology Detectable Cons gases Ref Micro cantilever Polymer (photoresist) VOC Power consuming [65, 66] Thermoelectric Polymer VOC Power consuming [67, 68] Semiconductor Metal oxide Power consuming [2, 23, 69] Optical Bacteria VOC Not reliable, response Chemiresistor Conductive polymer, etc. VOC Sensor drift [73-78] ChemFET Porous silicon/ metal oxide NO2, H2 Slow response [79-81] Sensor type 2.3.3 Sensing material CO, CH4 slow [70-72] Summary of integration approach for electrochemical gas sensor microsystem The comparison between hybrid integration and monolithic integration indicates that the monolithic approach leads to a cost-effective and highly-miniaturized microsystem. The 28 monolithic gas sensor literature review reveals to date no electrochemical gas sensor microsystem has been implemented by monolithic integration yet. Considering the dual benefits that the electrochemical sensing technique and monolithic approach bring, it is promising to develop an electrochemical gas sensor microsystem by monolithic integration. 29 3 Development of RTIL-based Sensor for Microsystem Platform 3.1 Overview RTIL-based electrochemical sensing technology was identified and chosen to develop sensor devices for wearable air pollutant monitoring microsystems due to its comprehensive strengths that are superior to other sensing technologies. However, traditional electrochemical sensor structure suffers slow response to gases. In this chapter, a new type of RTIL-based gas sensor structure is introduced to achieve rapid response time. Then, approaches for sensor miniaturization and sensor array packaging are reported. After that, the development of a flexible version of the RTIL-based gas sensor is presented. The achievements pave a path to realize a wearable gas sensor microsystem for air pollutant monitoring. 3.2 Development of fast-response RTILs-based electrochemical sensor 3.2.1 Sensor’s structure design for fast response RTIL-based electrochemical sensing technology inherits the outstanding performance of liquid-electrolyte electrochemical gas sensors, but also suffers slow response. The main obstacle to response time in existing RTIL-based electrochemical gas sensors is the slow diffusion of target gases from the RTIL surface to the electrodes. As illustrated in Figure 3.1(a), in the conventional sensor structure, the electrodes are fabricated on a substrate and RTIL is then coated on top of the 30 electrodes. Because the electrochemical reaction happens at the interface of the RTIL and the electrodes, a response cannot be measured until the gas analyte dissolves into the RTIL and reaches the electrodes. A full-scale response to gas analyte is achieved only after an equilibrium forms on the interface. The time interval between the gas reaching the electrode surface and establishing equilibrium is primarily dependent upon gas diffusivity in the RTIL. Due to RTIL’s high viscosity, RTIL-based electrochemical gas sensors suffer slow response caused by the low gas diffusivity in the RTIL. Different sensor structures have been developed to solve the issue of slow response. On the one hand, as response time is reciprocal to square of the thickness of electrolyte [82], it can be significantly reduced by thinning the RTIL layer [38, 45]. However, formation of thin RTIL layer is unstable and irreproducible [39]. By wicking RTIL in a permeable membrane, a controllable uniform electrolyte layer can be produced [35] (Figure 3.1(b)). However, the thickness of RTIL is limited by the thickness of membrane. On the other hand, the response time can be shortened by eliminating the gas diffusion process in RTIL. By placing electrodes between gaseous analyte and RTIL as shown in Figure 3.1(c) [37, 44], the response time can be significantly reduced. The drawback of this structure is that a very precise volume control is required to maximize the sensitivity. In contrast to the traditional structure, we have designed a planar-electrode-on-permeable-membrane (PEoPM) structure that bypasses the slow diffusion of gas across the RTIL. As illustrated in Figure 3.1(d), the electrodes are fabricated directly on a 31 Figure 3.1. (a) Conventional sensor structure: response time is slow due to slow gas diffusion through RTIL; (b) RTIL-in-permeable-membrane structure: response time is improved by a controllable thin electrolyte layer; (c) Porous-electrodes-on-RTIL structure: response time is improved by bypassing gas diffusion in RTIL. (d) Electrodes-on-permeable-membrane structure: response time is improved due to fast gas diffusion in the permeable membrane. gas-permeable membrane, allowing gas to reach the electrodes/RTIL interface through the permeable membrane, where diffusion is much faster than in the RTIL. If the porous membrane has hydrophobic response to RTILs while porous electrode has hydrophilic response to RTILs, RTIL can fill the porous electrodes but cannot soak into the porous membrane. In such case, the effective sensing area is only determined by the surface geometry of the porous electrodes rather than RTIL volume. Therefore, this solution can solve the volume control issue in Figure 3.1(c). 32 Choosing inert material to form the porous membrane, the sensor structure can be used to detect toxic gases. To simplify the sensor structure complexity and lower the fabrication cost, WE, CE and RE should be all fabricated on the permeable membrane. 3.2.2 Fabrication of PEoPM sensors Proper materials were selected for the PEoPM structure. POREX® porous PTFE with 35% porosity and 4μm pore size (Zitex TM, Chemplast, Incorporated, Wayne, New Jersey) was chosen as the permeable membrane due to its excellent inertness and hydrophobic response to RTIL. Gold was chosen as the electrode material because it is highly inert and can readily be deposited in thin films and patterned in a planar process. High-purity RTIL 1-butyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)imide ([C4mpy][NTf2]) was chosen as the electrolyte due to its low viscosity and high chemical stability. It was prepared and refined to gravimetric purities exceeding 99.9% using reported method [37, 44]. To verify the hydrophobic/hydrophilic property of blank porous PTFE membrane and gold-coated porous PTFE membrane, contact angles of an RTIL droplet on both membranes were measured using Video Control Angle System 2000. As shown in Figure 3.2, the contact angles are 95˚ and 18˚, respectively. The results demonstrate that the porous PTFE membrane has a hydrophobic response to RTILs while porous gold electrode has a hydrophilic response to RTILs. Therefore, the selected materials are capable of forming the proposed sensor structure. 33 Figure 3.2 (a) Image of RTIL [C4mpy][NTf2] droplet on porous PTFE. The contact angle is 95˚. (b) Image of RTIL [C4mpy][NTf2] droplet on gold-coated porous PTFE. The contact angle is 18˚. To realize the PEoPM structure, a fabrication process was developed. A gold film was deposited on a porous PTFE sheet using physical vapor deposition (PVD, Edward 360 thermal evaporator). Thicker films help to ensure continuity of the thin film electrodes applied to a porous substrate. Thinner films reduce blockage of PTFE pores, thus improving diffusion of gas to the WE/IL interface. This tradeoff was experimentally studied, and a gold film thickness of 300nm was chosen balance consistency and performance. Metal deposited on porous PTFE results in a very uneven surface that complicates patterning by traditional photoresist processes. To solve this problem, a stainless steel hard mask was prepared by electric discharge machining. The mask was tightly mounted against the PTFE substrate on the PVD sample stage, and after deposition it was removed to leave the desired electrode pattern. Finally, to add the RTIL layer, the electrodes were then coated with 200 µm-thick [C4mpy][NTf2] with a droplet process. The fabrication process is illustrated in Figure 3.3. 34 Figure 3.3. PEoPM sensor fabrication process. A stainless steel mask was mounted against the porous PTFE and then a gold film was deposited. After removing the mask, patterned planar gold electrodes were formed. Finally, RTIL was coated on the electrodes to form the sensor. Because RTIL is a highly resistive electrolyte (the conductivity of [C4mpy][NTf2] is only 0.22 S/m at room temperature [83] compared to 62 S/m for 50% H2SO4 [84]), the sensor electrodes were designed to minimize the electrolyte resistance and maximize the response current by interdigitating the WE and CE electrodes. An RE was included alongside the WE. The electrode structure shown in Figure 3.4 occupies 0.75” × 0.5” including large pads to facilitate 2 connections to instrumentation. The effective WE area is 10 mm . The finger width of and the gap between interdigitated CE and WE are both 200 µm. An O-ring was used to separate the sensing electrode area from the connection pad area during testing. 35 Figure 3.4. Photograph of fabricated PEoPM device. The WE and CE are interdigitated with finger widths of and the gaps of 200 µm. The O-ring isolates the sensor area from the contact pads. 3.2.3 Test of PEoEM sensors The fabricated sensor was placed in a gas chamber and sealed with an O-ring, as illustrated in Figure 3.5. To compare the response time of the different sensor structures shown in Figure 3.1(a) and Figure 3.1(d), both the upper chamber and lower chamber have a gas inlet and outlet for testing with gas flow. Path 1 over the top of the sensor corresponds to a conventional sensor structure while path 2 at bottom of the sensor corresponds to the PEoPM sensor structure. The total gas flow was maintained at 200 ± 0.05 sccm (standard cubic centimeters per minute) by digital mass-flow controllers (MKS Instruments Inc). EIS testing was carried out with a VersaStat MC potentiostat (Princeton Applied Research, Oak ridge, TN, U.S.A.). During EIS measurement, a sinusoidal AC signal with a 10mV peak-to-peak amplitude was applied. The DC bias was set to the electrochemical potential that 36 Figure 3.5. Cross-sectional view of the gas testing chamber. Separate inlets and outlets on the top and bottom of the chamber allow gas to be introduced along two separate paths to effectively characterize two different sensor structures, where path1 corresponds to a conventional sensor structure shown in Figure 3.1(a) and path 2 corresponds to the PEoPM sensor structure shown in Figure 3.1(d). generated the largest response for the chosen RTIL and the target gas. For the example pollutant gas SO2 reported below, the DC bias was set to -0.5V. To characterize the sensing response for the PEoPM device, SO2 was selected as an example pollutant. The sensor was exposed to SO2 and the sensor’s impedance spectrum was recorded. The AC stimulus signal was swept from 0.1 Hz to 10 kHz, and SO2 concentration was varied from 0 to 25 ppm. The recorded amplitude of the complex impedance |Z| is plotted in Figure 3.6., where each curve corresponds to the impedance spectrum under a fixed-concentration SO2. The RTIL sensor shows an obvious impedance amplitude change to different SO2 concentration, with a stronger response below 100 Hz. 37 To further illustrate the relation between the sensor impedance and SO2 concentration, the imaginary portion of the complex impedance Zim at 10 Hz were extracted to form an SO2 calibration curve. Figure 3.7 plots -Zim verses SO2 concentration. The experimental data was fit to an equation given by (see explanation from EIS model of RTIL-based gas sensor in Appendix A)  Z im  0.377  1.239CSO2 C SO2  7.588 (3.1) Figure 3.6. Impedance amplitude |Z| spectrum from 0.1 Hz to 10 kHz recorded from an RTIL-based PEoPM sensor presenting under a SO2 environment. SO2 concentration varies from 0 to 25ppm. Obvious impedance amplitude change to different SO2 concentration was observed, with a stronger response below 100Hz. 38 -Zim(kohm) 1.5 1 0.5 0 0 5 10 15 SO2 (ppm) 20 25 Figure 3.7. SO2 calibration curve for RTIL-based PEoPM sensor. Blue dots represent the recorded -Zim values at 10Hz. Red curve represents the fitting curve follows Equation 3.1. The 2 fitting curve achieves high least-squares correlation coefficients (R =0.9957). where CSO2 is the concentration of SO2 in ppm units. This fitting curve matches test data very well, 2 with least-squares correlation coefficients R =0.9957. Although the calibration curve is nonlinear, the sensitivity at any given concentration can be obtained by calculating the slope of the fitting curve (3.1) at the desired concentration. To compare the response times of the two different sensor structures shown in Figure 3.1(a) and Figure 3.1(d), 25 ppm SO2 was purged through path 1 and path 2 (Figure 3.5.) separately, and the transient response of impedance amplitude was recorded. Defining the baseline as 0% and the steady state as 100%, the data were normalized, and the transient response of the RTIL sensor is plot in Figure 3.8. For the conventional structure (corresponding to path 1), the RTIL sensor 39 reaches 90% of full-scale response (T90) in 56 s. However, using the PEoPM sensor structure (corresponding to path 2) the response time is reduced to only T90 = 20 s. The PEoPM sensor structure thus is 180% faster than the conventional structure in the case of SO2 sensing. These preliminary results demonstrate that an RTIL sensor on a PEoPM electrode structure using EIS methods provides a rapid and sensitive response to SO2. Experiments using other gas analytes such as CH4 show that similar rapid responses can be achieved. This new sensor structure thus appears well suited to measure multiple gases in a real-time air pollution monitoring system. Figure 3.8. Transient response of RTIL-PEoPM sensor to 25 ppm SO2, with gas flow from two different paths (see Figure 3.5.) representing a conventional sensor structure with 56 s full scale response and a new rapid-response structure that reduces response time to 20 s. 40 3.3 Miniaturization approach of PEoPM sensor arrays Preliminary results demonstrate that an RTIL-based sensor on a PEoPM structure provides a rapid and highly sensitive response to example air pollutants. To utilize the PEoPM structure in a wearable air pollutant monitoring microsystem, miniaturization of the sensor arrays is required. In this section, a miniaturization approach of the PEoPM sensor array is presented. 3.3.1 Fabrication of miniaturized PEoPM sensors The process described in the previous section successfully resulted a macro sensor device. However, the process is not applicable for miniaturized sensor fabrication. The stainless steel hard mask used in the macro sensor device fabrication was prepared by electric discharge machining only provides a resolution of around 200 µm. Thus the WE-CE gap could not be reduced below 200 µm, which required the electrodes to occupy a large area to keep the gap resistance reasonable. To overcome this significant limitation, a new fabrication process utilizing photolithography was developed. Microfabrication processes have been well developed and widely used in the semiconductor industry. Using photolithography followed by thermal metal deposition and liftoff process, electrode patterns on smooth substrates can easily be formed. Traditional photolithography has a resolution on the order of 1 µm and should be suitable for fabricating miniaturized gas sensors. However, directly applying traditional photolithography on porous PTFE substrates, as desired 41 here, introduces several problems that require development of a new photolithography process for porous substrates. First, note that photoresist is generally spin-coated onto a substrate that is vacuum-attached to the spinner. However, porous PTFE is soft and cannot be reliably held by vacuum because it is porous. A preliminary experiment demonstrated that, when a PTFE sheet was put directly on the spinner, significant photoresist would be sucked into the PTFE and could not be removed during liftoff. To resolve this problem, the porous PTFE was clamped to a glass substrate, and the glass was placed on the spinner and held by vacuum. Second, because porous PTFE has a rough surface, standard photoresist coatings of PTFE are not reliably smooth enough for patterning. For example, with porous PTFE having a 4 μm-pore size, around 4 μm of surface roughness can be expected. To resolve this problem, a thick film photoresist was selected. Hoechst AZ4620 was spin-coated at 2100 rpm to create 10 μm-thick layer. Lithography by 60 s UV exposure and 300 s developing in AZ300 MIF developer was found to reliably remove the thick resist. The profile scan by a Dektak3 Surface Profiler in Figure 3.9 demonstrates that the photoresist AZ4620 covered the rough surfaces of porous PTFE and that a good edge step, suitable for liftoff, was formed. Third, note that preliminary experiments found metal liftoff to be difficult, even after a gold-deposited sample was soaked in acetone for several days. One possible reason is that the photoresist became polymerized during the half hour PVD, where the temperature on the sample is over 100°C. Because polymerized photoresist cannot easily be dissolved in acetone, this would 42 Figure 3.9. Profile scan by Dektak3 Surface Profiler. The photoresist AZ4620 covered the cause the liftoff to fail. To solve this problem, note that exposure to UV should prevent photoresist from polymerizing. Thus, a 60 s flood exposure was performed after developing the photoresist and before metal deposition. Finally, because of the uneven distribution of pores on PTFE surface, the conductivity of metal traces were found to be unreliable and in the worst case could result in entire electrode fingers being disconnected, resulting in significant device-to-device variation. To resolve this problem, one possible solution is to increase the width of the electrodes thereby reducing the chance of a discontinuity. However, this approach conflicts with the goal to miniaturize the device. To analyze the relation between device variation and electrode width, 4 groups of 10mm-long gold traces with different widths were fabricated. Each group had eight identical elements, and resistance values were measured to compare their average sheet resistance and 43 standard deviation as shown in Table 3.1. As expected, the results show that traces with larger widths have better consistency. 200 µm width was chosen as a compromise between reliability and the miniaturization goal. By addressing each of the challenges described above, the RTIL-based gas sensor fabrication process was optimized to achieve a miniaturized PEoPM structure followed by the process addressed as shown in Figure 3.10. The porous PTFE sheet was clamped on glass wafer and spin-coated by AZ4620 photoresist. 5 min soft bake at 95°C was performed, followed by a 60 s UV exposure through the electrode pattern mask. After 300 s developing in AZ300 MIF developer, the sample was flood exposed for 60 s. Then a 300 nm-thick gold film was deposited on the porous PTFE sheet using physical vapor deposition (PVD, Edward 360 thermal evaporator). After deposition, the sample was soaked in acetone overnight and the gold electrodes pattern was formed by liftoff. Finally, to add the RTIL Table 3.1. Average sheet resistance and standard deviation with different width configurations. standard deviation (%) 0.288 standard deviation (Ω) 0.044 200 0.323 0.041 12.7 400 0.327 0.034 10.3 800 0.306 0.023 7.5 width (µm) average (Ω) 100 44 15.2 Figure 3.10. Miniaturized RTIL-based PEoPM sensor fabrication process. interface, the electrodes were then coated with 200 µm-thick [C4mpy][NTf2] with a droplet process. 45 Following the microfabrication process above, the miniaturized PEoPM device shown in Figure 3.11 was fabricated. The electrode structure occupies a 2 mm × 2 mm sensing area, only 8% of the area in the macro-scale device as shown in Figure 3.4. The WE and the CE were interdigitated for impedance measurement with a 200 µm width and a 100 µm gap. 3.3.2 Package of miniaturized PEoPM sensor arrays A miniaturized PEoPM device fabrication process was describe in previous section. To integrate the sensor within the wearable air pollutant monitoring system, the microfabricated PEoPM sensors should be packaged in an array form and electrically connected with sensor interface circuit. An individual RTIL-based PEoPM sensor element is a standard electrochemical cell that consists of three electrodes (WE, CE, and RE), and electrolyte. The three electrodes are on the PEoPM sheet and RTIL is the electrolyte. To form PEoPM sensor arrays, there are two schemes: Figure 3.11. Photograph of fabricated miniaturized PEoPM device. The WE and CE are interdigitated with finger gaps of 100µm. The O-ring isolates the RTIL area from the contact pads. 46 electrodes array that all sensor elements share the same RTIL or electrochemical cell array that each group of three-electrodes pairs with separate RTIL. The former scheme would be to directly pattern multiple electrodes on one piece of porous PTFE sheet with coating a continuing RTIL layer. However, this approach fails the sensors’ function according to the discussion below. The gas sensing selectivity is introduced by electrochemical biasing potential. Each sensor tailored to a specific gas target is biased at a specific potential. If the sensor elements in the array share the same electrolyte environment, the different electrical bias conditions would introduce significant interference among sensors. Therefore, electrochemical cell array scheme was chosen to develop PEoPM sensor arrays. To properly separate the electrochemical cells, O-rings were used to confine the RTIL in individual sensor element region. To avoid RTIL leaking from the edge between O-rings and porous PTFE, O-rings and PEoPM sheet were tightly clamped between two rigid boards. Both the up-board (lid) and lower-board (substrate) were designed containing a hole for gas diffusing to porous PTFE and for introducing RTIL to the electrochemical cell, respectively. In our design, Viton O-rings were chosen for its good seal and chemical resist performance; printing circuit board (PCB) was chosen to construct the lid and the substrate because it is mechanically rigid and easy for wire routing. The PEoPM sensor array were assembled as illustrated in Figure 3.12(a). First, four electrodes-patterned PEoPM sheets were placed on wire-routed PCB substrate, with pads connected to wire contacts on PCB by conductive epoxy. Then O-rings were attached on a windows-opened PCB lib and mounted on the wire-routed PCB substrate by tighten bolts, to create 47 RTIL reservoirs. Finally, [C4mpy][NTf2] were dropped in the reservoirs to create 200µm-thick RTIL sensing layer. Photos of the PEoPM sensor array are shown in Figure 3.12 (b). The sensor array occupies 1.55” x 1.25” including the connector to be electrically connected with sensor interface circuit. This array contains only four sensors because only four gas analytes were available for preliminary test. Nevertheless, using the package process described above, arrays contain eight sensor elements or more can be readily implemented. 3.3.3 Test of miniaturized PEoPM sensor arrays To demonstrate the functionality of the microfabricated PEoPM sensor array, the packaged Figure 3.12. (a) Illustration of the parts assembly for PEoPM sensor array. First, four electrodes-patterned PEoPM sheets were placed on wire-routed PCB substrate, with pads connected to wire contacts on PCB by conductive epoxy. Then O-rings were attached on a windows-opened PCB lib and mounted on the wire-routed PCB substrate by tighten bolts, to create RTIL reservoirs. Finally, [C4mpy][NTf2] were dropped in the reservoirs to create RTIL sensing layer. (b) Photographs of assembled PEoPM sensor array from front, back and side views. The sensor array occupies 1.55” x 1.25” including the connector to be electrically connected with sensor interface circuit. 48 sensor array was placed in a dessicant chamber and gas analytes were purged with gas flow maintained at 200 ± 0.05 sccm. CH4, SO2, NO2, and O2 were selected as example pollutants. CH4 concentration was varied from 0 to 5%; SO2 concentration was varied from 0 to 25ppm; NO2 concentration was varied from 0 to 35ppm; O2 concentration was varied from 10% to 20%. EIS testing was carried out with VersaStat MC potentiostat. The DC bias was set to the electrochemical potential that generated the largest response for the chosen RTIL and the target gas. For the example gas CH4, SO2, NO2, and O2, the DC bias of each channel was set at -0.3 V, -0.5 V, 1.2 V and -1.2 V to acquire the maximum response to target each gas analyte, respectively. The impedance spectra were recorded. Depending on the gas analyte being measured, the sensor’s impedance will change either due to a physical absorption of the analyte [85] or by chemical interaction with the analyte [46]. Physical absorption mechanisms will result in a change of the complex impedance (Z = Zre+jZim) at low frequency that will be most evident within the imaginary portion (Zim) and have little impact on the real portion (Zre). Chemical interaction mechanisms will results in a change of recorded current at low frequency that will be most evident within the amplitude of impedance (|Z|). For CH4 and SO2 sensing, given their physical absorption mechanisms, Zim at 10 Hz was extracted to plot calibration curves. For NO2 and O2 sensing, given their chemical reaction mechanisms, |Z| at 10 Hz was extracted to plot calibration curves. Figure 3.13 plots all the four gas analytes’ calibration curves for the RTIL sensor array. The red lines represent the fitting curves 49 with equation forms that are derived from the theory (see details in Appendix A). The impedance –Zim or |Z| of each gas sensor are expressed as follows  Z im ,CH  26.89  4 7.815CCH 4 CCH  1.622 4 2 (R =0.994) (3.2)  Z im, SO  22.54  2 10.95CSO 2 CSO  6.28 2 2 (R =0.986) (3.3) | Z |NO  2.05  101  [1.49  103  C NO  (4.72  102 +7.85  104  C NO ) j ]1 2 2 2 2 (R =0.967) (3.4) | Z |O  2.05  10 1  [8.51  10 3  CO  (2.92  10 2 +6.12  10 4  CO ) j ]1 2 2 2 (R =0.999) 2 (3.5) 50 35 32 -Zim(kohm) -Zim(kohm) 34 30 28 26 0 2 4 CH4 (%) 0 10 20 SO2 (ppm) 30 10 O2(%) 20 40 20 |Z|(kohm) |Z|(kohm) 25 20 6 25 15 10 30 0 10 20 30 NO2(ppm) 30 20 10 0 40 0 5 15 Figure 3.13. CH4, SO2, NO2, O2 calibration curves for RTIL sensors from EIS test data at 10Hz. where CCH4 is the concentration of CH4 in % units, CSO2 is the concentration of SO2 in ppm units, CNO2 is the concentration of NO2 in ppm units, and CO2 is the concentration of O2 in % units. 2 All least-squares correlation coefficients R are higher than 0.96, demonstrating the test results have good agreement with the theory. The results also demonstrate good sensitivity over the range of interest for the gas targets. To demonstrate the selectivity of the miniaturized sensor array, 20.4% O2, 3% CH4, 10 ppm SO2, and 10 ppm NO2 were purged into the test chamber sequentially in four steps and the sensor 51 array’s impedance responses were recorded in each step. Taking the impedance response to O2 as 100% signal S, the sensor array responses to all four gases are normalized as ΔS/S (%) and plotted in Figure 3.14. The appended gases have negligible interference to other gas sensors. This miniaturized sensor array is thus well suited to measure multiple gases in a wearable real-time gas monitoring microsystem. Figure 3.14. Selectivity experiment. 20.4% O2, 3% CH4, 10 ppm SO2, and 10 ppm NO2 were purged into the test chamber sequentially in four steps and the sensor array’s impedance responses were recorded in each step. Taking the impedance response to O2 as 100% signal S, the sensor array responses to all four gases are normalized as ΔS/S (%) and plotted. 52 3.3.4 Integration of PEoPM sensor array within a wearable monitoring system The miniaturized sensor array was successfully installed in a prototype of a wearable air pollutant monitoring system as shown in Figure 3.15. It consists of a packaged miniaturized sensor array, multi-electrochemical instrumentation board (MEIB), and a µ-Controller (µC) board. The size of the system is 2.6” × 2” × 1”. Details of the wearable system is described in [86].Development of the microsystem is still ongoing. 3.4 Development of flexible PEoPM sensors Recently, research on wearable sensors that features flexibility has drawn a lot of attention because physical flexibility of sensors could open a lot of new applications such as artificial skin, conformal monitoring, etc. Likewise, by featuring flexibility on gas sensors, the wearable air Figure 3.15. 2.6” × 2” × 1” prototype wearable air pollutant monitoring system with miniaturized sensor array, multi-electrochemical instrumentation board (MEIB), and a µ-Controller (µC) board. 53 pollutants monitoring microsystem could even be integrated into clothing. The previous section introduced a PEoPM structure that provides a fast response. However, the sensor’s package is mechanically stiff. The following section discusses how to design planar PEoPM sensors with mechanical flexibility. 3.4.1 Fabrication of flexible PEoPM sensors A PEoPM sensor array is mechanically stiff because two rigid boards are required to clamp the O-ring and PEoPM sheet to create reservoirs for RTILs. Therefore, flexible PEoPM sensor is possible only when the O-ring reservoir can be removed in the sensor design. However, without constrained by O-ring reservoir, thick RTIL (200µm in PEoPM sensor array) can easily flow out with small mechanical disturbance. Only thin RTIL layer can be formed on PEoPM sheet. Because RTIL has high viscosity, it would typically stick to the porous surface regardless mechanical disturbance. However, a thin layer cannot be formed by directly dropping pure RTIL on the PEoPM sheet. That’s because porous PTFE is hydrophobic to RTIL (Figure 3.2(a)), while gold on porous PTFE is hydrophilic to RTIL (Figure 3.2(b)). The surface energy difference for different substrates results in RTIL holding on individual electrodes as “islands” when applying a small volume (0.2 μL in the experiment) of pure RTIL on a PEoPM sheet, as shown in Figure 3.16 Case 1. If the RTIL islands become disconnected, the electrolyte fails to form an electrochemical cell 54 Figure 3.16. Coating experiments illustration. Case1 illustrates small volume pure RTIL coating (Vol.=0.2 μL) result. RTIL droplets were coated only electrodes surface. Case2 illustrates large volume pure RTIL coating (vol.=1 μL) result. RTIL droplets were coated on electrode surfaces as well as the gap between electrodes. Case3 illustrates ethanol/RTIL mixture coating (Rethanol/RTIL=2, RTIL’s vol.=0.2 μL) result. The initial mixture droplet was coated the electrodes area as well as the gap between electrodes. After ethanol removal, RTIL still occupied the initial droplet region. Case4 illustrates ethanol/RTIL mixture coating (Rethanol/RTIL=4, RTIL’s vol.=0.2 μL) result. The initial mixture droplet filled the pores inside the porous PTFE as well as covered electrodes surface. After ethanol removal, RTIL still stayed in the porous PTFE. over the electrodes. Although islands could merge into a bulky droplet when applying a large volume (1μL in the experiment) of pure RTIL, as shown in Figure 3.16 Case 2, this large droplet would easily flow away with small mechanical disturbances, ruining the sensor. 55 One approach is to mix RTIL with solvents that have low surface tension on porous PTFE. Ethanol was chosen for low surface tension on both porous PTFE and gold. It can be mixed with RTIL in any ratio. In addition, ethanol can easily be removed through evaporation. By fixing the RTIL volume to 0.2 μL, droplets with different ratios of ethanol to RTIL (denoted by Rethanol/RTIL) were dropped on the PEoPM surface, and the PEoPM sheet was then vacuumed for two hours to remove ethanol. These experiments are illustrated in Figure 3.16 Case 3 and Case 4. In Case 3, the initial mixture droplet coated not only the electrode area but also the gap between electrodes in the same electrochemical cell. After ethanol removal, RTIL still occupied the initial droplet region, and a conformal thin layer of RTIL was formed on both PTFE and electrodes surface. AC 2 impedance measurement between electrodes shows a resistance on the order of 10 Ω, implying good conductivity of electrolyte within the electrochemical cell. In Case 4 with higher ethanol concentration, the mixture’s surface tension was too small and the droplet filled the pores inside the porous PTFE as well as covering electrode surface. After ethanol removal, RTIL still remained within the porous PTFE. AC impedance measurement between electrodes shows an open circuit. 6 Even if the volume of the droplet was increased, a large resistance (on the order of 10 Ω) was measured for this ethanol/RTIL ratio due to uneven pore distribution and limited pore connection inside the porous PTFE. Following this procedure, optimized mixing ratio (Rethanol/RTIL=2) and droplet volume (VRTIL=0.2 μL) were experimental obtained to form a thin and conformal layer on the PEoPM surface that achieved a good, conductive electrochemical cell. 56 After the microfabrication (Figure 3.10) and droplet processes above, the PEoPM RTIL-based sensor shown in Figure 3.17(a) was produced. Concentric ring electrode geometry was adopted for amperometric-type planar sensors. A disk-shaped WE with radius of 1 mm was patterned in the center of porous PTFE sheet. An annular CE with inner radius of 1.125 mm and outer radius of 1.75 mm and an RE was patterned around the WE. The gap between WE and CE is 2 125µm. The total sensing region occupies 9.6 mm . A 21 µm-thick [C4mpy][NTf2] layer was coated on the sensing region, covering all electrodes and the gaps between them. The resulting sensor structure can be bent either convex or concave as shown in Figure 3.17, demonstrating its physical flexibility. Figure 3.17. (a) Photograph of microfabricated flexible PEoPM RTIL-based sensor. (b) Photograph of convex bend of the PEoPM RTIL-based sensor. (c) Photograph of concave bend of the PEoPM RTIL-based sensor. 57 3.4.2 Test of flexible PEoPM sensors Chemical experiments were carried out using the test setup as shown in Figure 3.15. O2 was selected as an example analyte. O2 and N2 (as background gas) concentrations were controlled using digital mass-flow controllers (MKS Instruments, Inc.) to vary O2 concentration from 0% to 21%. The total flow rate was fixed to 100 sccm to circumvent the influence of differential gas flow rate. The PEoPM RTIL-based sensor was placed in a flow cell within a desiccant gas chamber parallel to the gas flow as shown in Figure 3.18. The gas flows across both side of the sensor and reaches the electrodes both by permeating through porous PTFE and by dissolving into RTIL. Based on previously reported experiments [87], the response time is determined by the faster path permeating through the porous PTFE. Through wires that were attached to the sensor pads with conductive epoxy, the sensor was connected to a VersaStatMC potentiostat for electrochemical test. Figure 3.18. Gas test setup and position of the sensor within the flow cell. 58 To verify the functionality of the PEoPM RTIL-based sensor, cyclic voltammetry (CV) was performed between 0 - -1.8 V with 500 mV/s scan rate for the reduction of O2 at the WE, as shown in Figure 3.19 where the O2 concentration ranges from 0% to 21%. A broad current peak was observed at about -1.4V vs Au, which is in agreement with the CVs obtained for O2 reduction in RTIL reported in [88]. The relationship between peak current Ip(µA) and O2 concentration CO2(%) was found to fit the following linear calibration condition Figure 3.19. Cyclic voltammograms (scan rate = 500 mV/s) for the reduction of O2 at WE. O2 concentration varies from 0-21%. O2 reduction current peak was observed at -1.4V vs Au. The relationship between peak current and O2 concentration from 0% to 21% is linear. 59 2 Ip=-1.25 CO2 -3.35 (R =0.989) (3.6) This relationship demonstrates the functionality and linearity of the PEoPM RTIL-based sensor. For real-time monitoring, the constant-potential amperometry method is usually used due to its simple hardware implementation. To characterize the performance of the PEoPM RTIL-based sensor using constant-potential amperometry, the sensor was biased at the O2 reduction potential of -1.4 V and the current was measured. The sensor’s sensitivity, linearity, repeatability and limit of detection (LOD) were characterized as follows. The sensor’s response linearity was characterized by a step change of O2 concentration from 0% to 21% and then back to 0%. The response current vs. time curve shown in Figure 3.20 was recorded. The stable values at each step were collected to form the O2 calibration curve shown in Figure 3.21. The relationship between amperometric current Ia (µA) and O2 concentration CO2 (%) can be expressed by 2 Ia=-0.48 CO2 -0.64 (R =0.997) 60 (3.7) Figure 3.20. Current vs. time curve at various O2 concentrations when the potential is held at –1.4 V vs Au. N2 is the background gas. O2 concentration steps up from 0% to 21% and steps down from 21% to 0%. Figure 3.21. O2 calibration curve for RTIL sensor using data extracted from Figure 3.19. 61 2 The sensitivity of the amperometric response was measured as 0.48μA/%. The R value of 0.997 implies a good linearity of the sensor response to O2. Repeatability test was performed by alternatingly purging 1% O2 and pure N2. The resulting amperometric current is shown in Figure 3.22. Although the sensor shows a slight drift due to decay of the double layer capacitor charging current, by subtracting the baseline value the response current ΔI shows a good repeatability with an average value of 0.73 µA and a standard deviation of only 1.5 nA. The corresponding LOD for O2, defined as three times the baseline noise or 0.038 µA, was found to be 0.08 vol %. Figure 3.22. Constant-potential current response measured over three cycles of alternate exposure to 1% O2 and pure N2 flow at an applied potential of -1.4 V. 62 3.5 Conclusion This chapter introduced the implementation of robust RTIL-based electrochemical gas sensors featuring a planar-electrode-on-permeable-membrane structure that provides flexibility, high sensitivity and fast response. Fabrication processes have been developed for PEoPM sensor, miniaturized PEoPM sensor array, and flexible PEoPM sensor, respectively. Using SO2, CH4, NO2, and O2 as example analytes, the sensors’ functionality and performance were demonstrated. 63 4 Development of a Compact Low-Power Amperometric Instrumentation with Current-to-Digital Readout 4.1 Overview To develop wearable air pollutant monitoring microsystem, the sensor interface electronic circuits need optimized to minimize its power and complexity. For electrochemical sensors, amperometric technique is widely used. Miniaturization and optimization of amperometric instrumentation is essential for wearable applications such as gas sensor microsystems. The amperometric instrumentation consists of two parts: a potentiostat and a current readout circuit. The potentiostat provides current required for the reaction while maintaining the electrode/electrolyte interface at the correct potential. The current readout circuit conditions and digitizes the reaction current. Existing research focuses on optimizing individual parts (either potentiostat or readout circuit) given power/size/resolution requirements [47-49, 52, 56]. In contrast, an opportunity exists to optimize instrumentation circuit topology in terms of complexity from the perspective of circuit system level, or even the perspective of sensor system level. This chapter introduces a novel amperometric instrumentation circuit topology optimization from the perspective of sensor system level, resulting in a compact amperometric instrumentation circuit with current-to-digital readout. 64 4.2 Background 4.2.1 Electrochemical sensors and equivalent circuit models Electrochemical sensors in amperometric mode work under the following sensing principle: the reaction current is proportional to the analyte concentration when reacted electrode/electrolyte interface is biased at a constant voltage. To accurately control the reaction taking place at the interface, a three-electrode cell configuration has been applied to amperometric electrochemical sensors. In such three-electrode cell, the reaction takes place at the interface between the WE and electrolyte. A constant potential is maintained between the RE and the WE. The CE provides a current path to the WE. To analyze the electrochemical sensor’s electrical response, equivalent circuit models have been proposed in EIS theory. Randles circuit model [46] shown in Figure 4.1 (a) is a classic equivalent circuit model widely used to describe a three-electrode sensor. The impedance between RE and WE consists of an uncompensated solution resistor Rs (relatively small), in series with the parallel combination of the double layer capacitor Cdl at the WE interface (charging current ic follows through this path), and an impedance of a Faradaic reaction caused by AC stimulus (AC Faradaic current if follows through this path). The Faradaic reaction consists of a charge transfer resistor Rct and Warburg element Zw ZW  65 AW j (4.1) Figure 4.1. Equivalent circuit model of electrochemical sensor cell. (a) Randles model (b) Complete model considering both AC and DC stimulus. (c) Simplified model for circuit analysis. where Aw is Warburg coefficient and ω is the angular frequency. Since our only interest is at the WE interface, the impedance between the CE and the RE is denominated as simple impedance Z. Notice that this model only represents sensor’s response to small AC stimulus. To represent both AC and DC response, a complete equivalent circuit model is shown in Figure 4.1(b) [89, 90]. A 66 current source is added to represent DC Faradaic current If. Here, If is the constant reaction current proportional to the analyte concentration in amperometric electrochemical sensors, which is the main interest in sensor current measurements. In general, if << If, and Rs is relative small. They can be considered as second-order effects in sensors response. For analysis simplicity, Rs and if are initially omitted in the following circuit topology derivation, and the negligible impact of this omission will be discussed in Section 4.4. The simplified model is shown in Figure 4.1(c). 4.2.2 Traditional amperometric instrumentation An amperometric instrumentation circuit consists of two parts: a potentiostat and a current readout circuit. The potentiostat provides current from the CE to the WE while maintaining the voltage between RE and WE. A typical potentiostat can be implemented by a single opamp with appropriate connections [89, 90]: the positive input node is connected with bias for RE VRE, the negative input node is connected with RE, and the output is connected with CE to provide current. The current readout circuit collects If either at WE or CE, then conditions and digitizes it. Two topologies are used to implement the current readout circuit: a current-to-voltage convertor followed by a voltage-mode analog-to-digital convertor (ADC) [49, 91-93], or a single current-mode ADC [52, 94]. Given its compact structure, the current-mode ADC is usually chosen for readout circuit design. 67 Given the requirement of sensor applications for low power and low complexity, a traditional amperometric instrumentation circuit utilizes the single opamp for potentiostat design and the current-mode ADC [94] for current readout circuit design. The schematic is shown in Figure 4.2. The circuit working principle is described as follows. Two reference current sources Iref with opposite directions alternatively connect with the integrator through switches, which are controlled by the digital output of the hysteresis comparator Dn. Thus the input current of the integrator Iint is I int  I f  (1) Dn  I ref (4.2) As the waveforms in Figure 4.3 illustrate, the integrator’s capacitor is charged/discharged according to the direction of Iint. Consequently, the output of the integrator Vint rises/falls correspondingly to Iint direction. While Vint reaches the hysteresis comparator upper/lower bound Figure 4.2. Schematic of a traditional amperometric instrumentation circuit including potentiostat and current-mode ADC. 68 Figure 4.3. Waveforms of the current on the integrator input Iint, the voltage on the integrator output Vint, and the digital output of the comparator Dn. (Vref+/-ΔV/2) (where ΔV is the hysteresis window width and Vref is the reference voltage), Dn flips, changing Iint following (4.2). The square waveform at the output of the hysteresis comparator is then digitized by a counter with the reference clock at a much higher frequency. The time interval of the digital “high” for Dn is T1  Cint  V I ref  I f and the time interval of the digital “low” for Dn is 69 (4.3) Cint  V I ref  I f T0  (4.4) From (4.3) and (4.4), If can be expressed as a function of Iref, T1, T0 If  T0  T1 I ref T0  T1 (4.5) Defining the duty cycle α as  T1 T1  T0 (4.6) Substituting T1 and T0 by α in (4.5), If can be expressed as a function of α and Iref I f  (1  2 )  I ref (4.7) Therefore, given a known Iref, If is obtained by measuring duty cycle of Dn. Notice that If is independent of both the integrator capacitor Cint and the hysteresis comparator parameters (ΔV and Vref). 4.3 Compact amperometric instrumentation design The previous section introduced the electrochemical sensor equivalent circuit model and the typical amperometric instrumentation circuit. Substituting the sensor symbol in Figure 4.2 by the simplified sensor equivalent circuit model (Figure 4.1(c)), the schematic of the electrochemical 70 Figure 4.4. Schematic of the electrochemical sensor system. It consists of the traditional amperometric instrumentation circuit and the simplified electrochemical sensor equivalent circuit model. sensor system is given in Figure 4.4. Notice that the sensor operates at the steady state when no current flows through Cdl, only If is collected in the readout circuit. From systematic point of view, the sensor system contains two capacitors: Cdl and Cint. Cint is part of the readout circuit and used for charging/discharging; Cdl is the inherent interface capacitor, but doesn’t play a functional role in the system. Since capacitors usually occupy large area in a single IC chip, if Cdl can play the role of Cint, Cint can be eliminated to save area. How to modify the circuit to incorporate Cdl into the amperometric instrumentation? An answer is given in this section by deriving a compact amperometric instrumentation topology from the traditional structure. Part of the sensor model and the current-readout circuit are taken as starting point for the derivation. Ignoring sensor’s bias requirement at this step, a current source If is used to substitute the sensor model. As shown in Figure 4.5, given node B is a low-impedance node, folding the current source to the output of the integrator is equivalent to the typical topology of the current 71 Figure 4.5. Derivation of the instrumentation topology. The input current source is folded to parallel connect with integrator capacitor. readout circuit. (4.7) holds given a satisfied sensor’s bias condition. Notice that the parallel connection of If and Cint is the same as the equivalent circuit between RE and WE as shown in Figure 4.1(c) and Cint can be arbitrary, (4.7) is still valid when Cint is substituted by Cdl. To satisfy sensor’s bias condition, a potentiostat function is incorporated into the current-mode ADC by the following modification steps. First, by flipping the direction of If and substituting Vref with VWE and VWE with VRE, the voltage between the RE and the WE can be held by feedback loops of the integrator (loop1) and of the ADC (loop2). Although the WE potential is not strictly held constant due to a nonzero value of ΔV in the loop2, the perturbation on the WE does not affect the sensor’s steady state as long as ΔV is set small enough (less than 10 mV) [95]. In addition, because current can only flow from the CE to the WE, node A should be connected to the CE rather than the RE. 72 Following the modification steps described above, a modified amperometric instrumentation circuit with the sensor model is illustrated in Figure 4.6. Because the direction of the current source If is opposite from the direction of If in Figure 4.5, If in Figure 4.6 is written as I f  (2  1)  I ref (4.8) This topology successfully realizes the functions of both current-mode ADC and potentiostat. Compared to a traditional topology in Figure 4.2, it utilizes Cdl for integrator, and eliminates one opamp required by the potentiostat and Cint required by the integrator. Additional optimization is conducted as follows. Notice that voltages at the RE and the WE are both held by the feedback loops and no constrains are required for VRE and VWE from the circuit perspective. Therefore nodes RE and WE are interchangeable. By swapping WE with RE, an optimized structure is shown in Figure 4.7. This topology is equivalent to the topology in Figure 4.6. Thus If follows (4.8). Notice that the WE is connected with a unit-gain buffer. This buffer can Figure 4.6. Schematic of the modified amperometric instrumentation circuit with sensor equivalent circuit model. 73 be discarded for further optimization. By connecting the WE with ground and substituting VRE with VRE-WE, a compact current-to-digital amperometric instrumentation (CCDAI) topology is given in Figure 4.8 as equivalent to the optimized structure in Figure 4.7. Here, we assume that sensor’s bias requires VRE>VWE, concluding VRE-WE>0. If sensor’s bias requires VRE