TFT’s Circuit Simulation Models and Analogue Building Block Designs Xiang Cheng Department of Engineering University of Cambridge This dissertation is submitted for the degree of Doctor of Philosophy Homerton College February 2018 I would like to dedicate this thesis to my loving parents and my beautiful and considerate wife . . . Acknowledgements I would like to express my deep sense of gratitude to my supervisor Prof. Arokia Nathan, who gave me the opportunity to study in one of the most known universities in the world. He is not only a brilliant supervisor who provides sufficient guidance, knowledge and care, but also a very nice person who treats students like family members. The dissertation could not be finished without his help especially in its scope and vision. The flexible group culture gives us sufficient time and space to think deeply about research and life, which I surely benefit a lot from. Secondly, I would also like to thank my parents for their caring and support during the years. As the single child of the family, I am always overwhelmed by the care and support they have given to me. Now it is approaching the end of my PhD study and I have started my own family. Living on my own does make me understand more of your words and the emotions under your tongue, this could only make me feel more appreciated for your care and support. I would like to thank my wife Rui Mo. A lot has happened between us and our lives during this 4 years. We were boyfriend and girlfriend and now we are husband and wife. We were both students when we arrive in UK and now we are both doctors. It is very fortunate for me to have you around my life. I cannot finish the dissertation at least to the same quality without your support and I certainly would not live as well. Many thanks to Dr. Sungsik Lee, whom I worked close together with. It has been a very nice time chatting, drinking and smoking (although I only smoke second-hand) together. He has given me many valuable advice in research from the very detail like drawing figures to the general question as what is a valuable research. A great portion of this dissertation has been influenced by him. More importantly, the discussions has helped me a lot regarding how to conduct research and how to distinguish the value of research from industry. I would like to thank Shuo Gao for the accompany of sitting next to me through out the 4 years study. It’s been great discussing with him mostly everything about life and research. Although we were working on different topics, the general discussions really keeps me thinking. Also his productiveness also pushes me to publish papers more efficiently in these years. vi Thanks to Jiahao Li and Dr. Hanbin Ma for the supervision and help on the bio-sensor project. Although it is a separate project from this thesis, it broadens my knowledge on the bio-sensing applications. Believe it or not, bio-sensing was intended to be one application example of this thesis. However, I later decided to delete the part due to the very different technology it was used in the bio-sensing project. Thanks to Guangyu Yao and Chen Jiang for doing the processing in the group. The hard working of both of you always pushes me forward. Thanks to Constantinos Tsangarides for the discussions for several side projects and the trust of me answering several of your difficult questions. I would also like to thank him for his patient explanation of religious stories which, as a Chinese, I’ve never heard much of. Thanks to Mojtaba Bagheri for the discussion over the AMOLED pixel project. There were several points that you have taught me which I would otherwise not know. Many thanks to Cambridge Commonwealth, European & International Trust and Chinese Scholarship Council for providing the scholarship. Abstract Building functional thin-film-transistor (TFT) circuits is crucial for applications such as wearable, implantable and transparent electronics. Therefore, developing a compact model of an emerging semiconductor material for accurate circuit simulation is the most fundamental requirement for circuit design. Further, unique analogue building blocks are needed due to the specific properties and non-idealities of TFTs. This dissertation reviews the major developments in thin-film transistor (TFT) modelling for the computer-aided design (CAD) and simulation of circuits and systems. Following the progress in recent years on oxide TFTs, we have successfully developed a Verilog-AMS model called the CAMCAS model, which supports computer-aided circuit simulation of oxide-TFTs, with the potential to be extended to other types of TFT technology families. For analogue applications, an accurate small signal model for thin film transistors (TFTs) is presented taking into account non-idealities such as contact resistance, parasitic capaci- tance, and threshold voltage shift to exhibit higher accuracy in comparison with the adapted CMOS model. The model is used to extract the zeros and poles of the frequency response in analogue circuits. In particular, we consider the importance of device-circuit interactions (DCI) when designing thin film transistor circuits and systems and subsequently examine temperature- and process-induced variations and propose a way to evaluate the maximum achievable intrinsic performance of the TFT. This is aimed at determining when DCI becomes crucial for a specific application. Compensation methods are reviewed to show examples of how DCI is considered in the design of AMOLED displays. Based on these design considerations, analogue building blocks including voltage and current references and differential amplifier stages have been designed to expand the analogue library specifically for TFT circuit design. The VT shift problem has been compensated based on unique circuit structures. For a future generation of application, where ultra low power consumption is a critical requirement, we investigate the TFT’s subthreshold operation through examining several figures of merit including intrinsic gain (Ai), transconductance efficiency (gm/IDS) and cut-off frequency ( fT ). Here, we consider design sensitivity for biasing circuitry and the impact of device variations on low power circuit behaviour. Table of Contents List of Figures xiii List of Tables xvii Nomenclature xix 1 Introduction 1 1.1 Thin-Film Transistors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.2 TFT Enabled Flexible and Transparent Electronics . . . . . . . . . . . . . 3 1.3 Research Goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.4 Thesis Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 2 TFT Compact Modelling 7 2.1 Computer Aided Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.1.1 Physical and Empirical Modelling . . . . . . . . . . . . . . . . . . 8 2.2 Cambridge’s TFT Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 2.2.1 Above-Threshold Model . . . . . . . . . . . . . . . . . . . . . . . 9 2.2.2 Subthreshold Model . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.3 Off Region Model . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.2.4 Unified Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.3 Computer Interpretation of Device Model . . . . . . . . . . . . . . . . . . 20 2.3.1 SPICE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 2.3.2 Verilog . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.3.3 Short Summary of Comparison . . . . . . . . . . . . . . . . . . . 24 2.4 CAMCAS Model in Simulation Environment . . . . . . . . . . . . . . . . 25 2.5 Summary and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 3 Small Signal Modelling 29 3.1 General Small Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 x Table of Contents 3.1.1 Midband Small Signal Model . . . . . . . . . . . . . . . . . . . . 29 3.1.2 High-Frequency Small Signal Model . . . . . . . . . . . . . . . . 31 3.2 TFT Small Signal Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 3.2.1 Midband TFT Model . . . . . . . . . . . . . . . . . . . . . . . . . 33 3.2.2 High-Frequency TFT Model . . . . . . . . . . . . . . . . . . . . . 38 3.3 Model Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 3.3.1 Measurement Setup . . . . . . . . . . . . . . . . . . . . . . . . . . 41 3.3.2 s-parameters and fT Measurement . . . . . . . . . . . . . . . . . . 42 3.3.3 fT and VT shift . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 3.4 Summary and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 4 TFT analogue circuit building blocks 63 4.1 Status of TFT Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 4.2 Challenges of TFT Analogue Circuits . . . . . . . . . . . . . . . . . . . . 63 4.3 Building Blocks and Simulations . . . . . . . . . . . . . . . . . . . . . . . 65 4.3.1 Voltage Reference . . . . . . . . . . . . . . . . . . . . . . . . . . 65 4.3.2 Current Mirror . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 4.3.3 Current Reference . . . . . . . . . . . . . . . . . . . . . . . . . . 69 4.3.4 Differential Amplifier . . . . . . . . . . . . . . . . . . . . . . . . 76 4.4 Full Op-Amp Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 4.5 Current Mirror Layout and Measurement . . . . . . . . . . . . . . . . . . . 78 4.6 Summary and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 5 Conclusion and Future Work 85 References 87 Appendix A Device Circuit Interaction 99 A.1 Importance of DCI in TFTs . . . . . . . . . . . . . . . . . . . . . . . . . . 99 A.2 Impact of TFT Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 A.2.1 Temperature Dependence . . . . . . . . . . . . . . . . . . . . . . . 100 A.2.2 Geometric Dependence . . . . . . . . . . . . . . . . . . . . . . . . 104 A.3 Compensation Methods in Circuits & Applications . . . . . . . . . . . . . 107 A.3.1 On-Pixel Programming . . . . . . . . . . . . . . . . . . . . . . . . 108 A.3.2 Off-Pixel Feedback . . . . . . . . . . . . . . . . . . . . . . . . . . 108 A.4 Specific Device Properties and Alternate Circuit Architecture . . . . . . . . 112 A.4.1 Analogue Gain Stage with Mono-type TFTs . . . . . . . . . . . . . 112 Table of Contents xi A.4.2 Persistent Photoconductivity . . . . . . . . . . . . . . . . . . . . . 115 A.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 Appendix B Subthreshold Operation for Schottky Barrier TFTs 119 B.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 B.2 Subthreshold Model for IGZO TFTs . . . . . . . . . . . . . . . . . . . . . 120 B.2.1 DC Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 B.2.2 Small Signal Model . . . . . . . . . . . . . . . . . . . . . . . . . 120 B.3 Figures of Merit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 B.3.1 Intrinsic Gain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 B.3.2 Transconductance Efficiency . . . . . . . . . . . . . . . . . . . . . 122 B.3.3 Cut-off frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 B.4 Sensitivity to Variations and Bias . . . . . . . . . . . . . . . . . . . . . . . 125 B.4.1 Current Sensitivity to Variations . . . . . . . . . . . . . . . . . . . 125 B.4.2 Accuracy Requirement for Biasing Circuitry . . . . . . . . . . . . 127 B.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 List of Figures 1.1 Conceptual structure of a bottom-gate TFT. . . . . . . . . . . . . . . . . . 2 1.2 Field effect mobility and stage delay for different semiconductor families. Captured from [? ]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 1.3 Flexible circuits and displays with TFTs (a) stress sensor array with built-in amplifiers (b) Σ−∆ ADC (c) op-amp (d-f) flexible and transparent displays. Figures are captured from [? ? ? ? ? ? ]. . . . . . . . . . . . . . . . . . . . 4 2.1 Carrier transport combining percolation with trap-limited conduction (TLC) for oxide semiconductor TFTs (Here, DB and WB denote spatial distance and width of potential barriers, respectively) . . . . . . . . . . . . . . . . . . . 9 2.2 Comparison between measured and modelled transfer characteristics (IDS vs. VGS) for the above-threshold region. For each region, a good agreement between measurement and modelling is achieved. . . . . . . . . . . . . . . 12 2.3 Comparison between measured and modelled transfer characteristics at dif- ferent temperatures. Here, the proposed model shows a better agreement compared to the conventional model (e.g. trap-limited conduction model). . 13 2.4 Density of states profile with (a) Fermi level within interface states at low gate voltage, and (b) the case of deep states dominant with increasing gate voltage corresponding Fermi level now within deep states. . . . . . . . . . 14 2.5 Measured and modelled subthreshold current (Isub) as a function of VGS . . 15 2.6 Measured and modeled IDS vs. VGS at subthreshold region for a different VDS. Inset: Measured and modeled IDS vs. VDS for VGS = 0.5V . Here, the equation is simplified from Eq. (2.5) introducing the pre-constant I0 which is around 10pA at VGS = 0.5V. . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.7 (a) Calculated IDS vs. VX of the combined above- and sub-threshold model for different VG (4, 6, 8V). (b) First, (c) second, (d) third, and (e) fourth deriva- tives of IDS with respect to VX . The inset of (a): test circuit configuration of the GST. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 xiv List of Figures 2.8 (a) Schematic IDS vs. VGS curves (gray circles) with the conventional power- law models for subthreshold (Sub-T, solid line) and above-threshold (Above- T, dot line) characteristics. In this case, we need two models for these two different operational regions. Here, Von and VT are on-voltage and threshold voltage, respectively. (b) Schematic IDS vs. VGS curves (gray circles) with only one model: unified model which can cover subthreshold as well as above-threshold regions at the same time. . . . . . . . . . . . . . . . . . . 19 2.9 Parameter extraction for unified model & measured and modelled transfer curve with percentage error . . . . . . . . . . . . . . . . . . . . . . . . . . 21 2.10 Verilog-A behaviour model in early mixed signal simulation environment . 23 2.11 Verilog-A(MS) extension for compact modelling with SPICE like simulator integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 2.12 Circuit design and simulation flow using CAMCAS model in Cadence Design Environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 2.13 A voltage amplifier used in simulation with CAMCAS model . . . . . . . . 26 2.14 Simulation results of the voltage amplifier . . . . . . . . . . . . . . . . . . 27 3.1 Small signal model for a general 3 terminal device . . . . . . . . . . . . . . 31 3.2 High frequency small signal models for CMOS and BJT . . . . . . . . . . 32 3.3 The inner structure based on which static TFT models are developed . . . . 34 3.4 Midband small signal models . . . . . . . . . . . . . . . . . . . . . . . . 35 3.5 Midband TFT model considering VT shift . . . . . . . . . . . . . . . . . . 38 3.6 Bottom-gate TFT structure considering passive components of the small signal equivalent circuit. COV S and COV D are the overlap capacitance at source and drain side; Cch is the channel capacitance; RS and RD are the contact resistances at source and drain side. . . . . . . . . . . . . . . . . . 39 3.7 High frequency small signal models . . . . . . . . . . . . . . . . . . . . . 39 3.8 Measurement setup for fT and s-parameters . . . . . . . . . . . . . . . . . 42 3.9 I-V characteristic of the TFT under test . . . . . . . . . . . . . . . . . . . 43 3.10 Comparison between terms of (1−S11)(1+S22) and S12S21 . . . . . . . . 45 3.11 Comparison between Eq. (3.20) and its approximation Eq. (3.21) . . . . . . 46 3.12 Equivalent circuit for S11 and S21 measurement . . . . . . . . . . . . . . . 47 3.13 Magnitude and phase measurement and simulation for S11 parameter based on TFT model and adapted CMOS model . . . . . . . . . . . . . . . . . . 48 3.14 S11 simulation of the TFT and CMOS models . . . . . . . . . . . . . . . . 50 3.15 Magnitude and phase measurement and simulation for S21 parameter based on TFT model and adapted CMOS model . . . . . . . . . . . . . . . . . . 52 List of Figures xv 3.16 S11 simulation of the TFT and CMOS models . . . . . . . . . . . . . . . . 54 3.17 Magnitude and phase measurement and fitting for S11 parameter based on single pole and zero approximation . . . . . . . . . . . . . . . . . . . . . . 57 3.18 h21 results with and without S11 fitting . . . . . . . . . . . . . . . . . . . . 58 3.19 S21 measurement showing low frequency range for gm extraction . . . . . . 59 3.20 fT vs. gm relation through S-parameter measurement . . . . . . . . . . . . 60 3.21 gm vs. VT shift relation through stress measurement with I-V sweep . . . . 60 3.22 Extracted ∆ fT vs. VT shift relation. The error bar here is the standard deviation of the fitting to obtain the corresponding result. . . . . . . . . . . 61 4.1 Voltage reference circuit with all TFTs working in saturation region . . . . 66 4.2 Simulation results for the proposed voltage reference circuit with two differ- ent values of β1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 4.3 Pseudo p-type current designed with all n-type TFTs . . . . . . . . . . . . 68 4.4 Simulation results of the proposed pseudo p-type current mirror . . . . . . 70 4.5 Fully n-type current reference circuit with pseudo p-type current mirror and a reference resistor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 4.6 Simulation results of the current reference circuit with resistor where W3,4 are simulated with different value to exhibt the change of working range . . 72 4.7 Fully n-type current reference circuit without resistor . . . . . . . . . . . . 73 4.8 Simulation results of the current reference circuit without resistor where WR is simulated with different value to exhibt the change of output current . . . 74 4.9 Improved current reference circuit with negative feedback on TR . . . . . . 75 4.10 Simulation results of the improved current reference circuit where WR is simulated with different value to exhibt the change of output current . . . . 75 4.11 Differential amplifier stage with pseudo p-type mirror . . . . . . . . . . . . 76 4.12 Simulation results of the proposed differential amplifier stage . . . . . . . . 77 4.13 Full op-amp design with supply independent bias . . . . . . . . . . . . . . 78 4.14 Simulation results of the proposed op-amp . . . . . . . . . . . . . . . . . . 79 4.15 Schematic of the fabricated ISO current mirror. . . . . . . . . . . . . . . . 80 4.16 Picture and measurement result of the fabricated current mirror based on a separate-device layout. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 4.17 Picture and measurement result of the fabricated current mirror based on a interdigitated stacked layout. . . . . . . . . . . . . . . . . . . . . . . . . . 82 A.1 Illustration of DCI in relation to performance requirements of a desired application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 xvi List of Figures A.2 I-V characteristics of the examined TFT under different temperatures . . . . 101 A.3 Extracted values of K, αp and VT at the different temperatures . . . . . . . 103 A.4 Normalized temperature sensitivity of current and the contribution of differ- ent parameters at 313K . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 A.5 Fitted probability distribution and histogram of the three key parameters. The data was extracted at room temperature from a 1080*1920 RGBW AMOLED panel with pixel circuits as described in Fig. A.7. . . . . . . . . . . . . . . 106 A.6 Normalized standard deviation of current and the contribution of different parameters(K, VT and α) . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 A.7 AMOLED pixel structure for off-pixel defect extraction and feedback . . . 109 A.8 Equivalent circuit for the defect extraction phase of OLED and TFT . . . . 110 A.9 Extracted aging parameters for (a) TFTs and (b) OLEDs after continuously displaying a checker board (W: displayed with white squares, B: displayed with black squares) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 A.10 Extracted aging parameters for (a) TFTs and (b) OLEDs after continuously displaying a checker board (W: displayed with white squares, B: displayed with black squares) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 A.11 (a) Common-source gain stage of an amplifier (b) sketch of different type of loads passing through the same bias point . . . . . . . . . . . . . . . . . . 113 A.12 Gain stage of an amplifier with (a) feedback load (b) boost strap load . . . . 114 A.13 Band diagram to explain the effect of positive gate pulse on recovery. . . . 116 A.14 Persistent photoconductivity (PPC) and its removal . . . . . . . . . . . . . 117 B.1 Simulation result for the intrinsic gain (Ai) of the Schottky-Barrier IGZO TFT (SB-TFT). The parameter n is the ideality factor of the source-semiconductor junction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 B.2 gm/IDS from deep subthreshold to above threshold region. The value reaches maximum at deep subthreshold and increases with steeper SS. . . . . . . . 123 B.3 Cut-off frequency vs. voltage bias for different SS. . . . . . . . . . . . . . 124 B.4 Normalized sensitivity to VT shift for different SS . . . . . . . . . . . . . . 125 B.5 Normalized sensitivity to SS variations for different SS mean values . . . . 126 B.6 The useful bias range for a depletion load common-source amplifier. (a) con- ceptual figure of the load line and transfer curve of the amplifier, where green curve illustrates the correct bias condition and blue curves illustrates that the amplifying TFT is biased out of useful bias range (b) circuit schematic of the amplifier (c) simulated useful bias range with respect to SS . . . . . . . 127 List of Tables 2.1 Extracted Subthreshold Parameters at T=300K . . . . . . . . . . . . . . . . 16 2.2 Extracted Parameters for Unified Model at T=300K . . . . . . . . . . . . . 20 2.3 Comparison between different model languages . . . . . . . . . . . . . . . 24 3.1 Summary of parameters for the CMOS and TFT models . . . . . . . . . . . 62 Chapter 1 Introduction The wide usage of amorphous silicon (a-Si:H) thin film transistors (TFT) in TFT-LCD display [? ] which dates back to the 1980s has boosted research on materials and the processing technology of TFTs [? ? ], and has subsequently enables another success in organic displays [? ]. Further improvements in TFT device performances and its flexibility, transparency, roll- to-roll compatibility and low cost make them promising in many other applications such as radio-frequency identification (RFID) tags [? ], sensors [? ] and wearable devices. In order to reduce the integration complexity and system cost, a system-on-panel approach has been pointed out aiming at producing all kinds of circuit building blocks on a single substrate [? ]. Although circuit blocks such as logic gates [? ], amplifiers [? ] and analogue-to-digital coverters (ADCs) [? ] have been reported in recent years, the capability of TFT circuits is still low. Complex systems with only TFTs are still lacking, especially for TFT analogue circuits and applications. The reasons are various. Firstly, due to the ever-changing nature of the semiconductors used in newly developed TFT technologies, device models can vary because of changes in the underlying physics. However, an accurate model for newly developed technology is essential for circuit design. Secondly, the design flexibility of TFTs is not comparable with the well-developed silicon metal–oxide–semiconductor field-effect transistors (MOSFET) or even bipolar junction transistors (BJTs). Lacking complementary semiconductors and the capability of making enhancement and depletion TFTs on the same substrate, many existing circuit blocks are not directly usable thus limiting the flexibility of designing TFT circuits. Therefore, the dissertation as a whole is dedicated to exploring and synthesising a more accurate model specifically for circuit simulation along with development of essential building blocks for analogue circuits so as to build a bridge between the existing circuit topologies to the TFT domain. 2 Introduction Fig. 1.1 Conceptual structure of a bottom-gate TFT. 1.1 Thin-Film Transistors It is surprising to find out that the first concept of the thin film transistor dates back to the 1930s, even before the first point contact transistor was demonstrated. The early failure of the attempt to make TFT devices and the later success of the crystalline silicon devices lead to an abandonment of the TFT concept in the industry until a killer application was identified, i.e. the TFT-LCD, in the 1980s [? ]. Extensive research on different materials started from then on. The TFT by definition is “a field effect transistor made of nonsingle crystal semicon- ductor film deposited on an insulating substrate”, quoting the definition used in [? ]. A conceptual figure of a TFT structure is shown in Fig. 1.1 where a bottom-gate structure is used. The device is designed following the same principle as other field effect transis- tors. The rule of thumb is that under working conditions, the vertical electric field through semiconductor is much higher than the horizontal electric field along the surface of the semi- conductor. Therefore, the gate electrode can have dominating control over the semiconductor characteristics. By virtue of its working principle, a TFT is the same as a floating body MOSFET. How- ever, as semiconductors used in these devices are normally amorphous or poly-crystalline, the effective mobility if much lower than single crystal silicon in MOSFETs, yielding a much lower operating current. This motivates researchers to improve the effective mobility of TFTs in subsequent years. 1.2 TFT Enabled Flexible and Transparent Electronics 3 Fig. 1.2 Field effect mobility and stage delay for different semiconductor families. Captured from [? ]. Up until now, many alternative materials have been found to exhibit higher or comparable effective mobility against the first widely used TFT technology which is based on a-Si. Fig. 1.2 lists the mobility achieved in a range of different semiconductor materials [? ]. Among the materials listed in Fig. 1.2, metal-oxide semiconductors stand out as the potential leading semiconductor material for TFTs as they exhibit high mobility and good uniformity in large scale and amenable to low-temperature fabrication. In addition, metal oxide semiconductors are very promising for transparent and flexible electronics due to their wide bandgap. Therefore, it will be used as the main consideration in developing the models and circuit blocks in this dissertation. 1.2 TFT Enabled Flexible and Transparent Electronics The low processing temperature of TFT technologies such as a-Si:H, metal oxide and organic TFTs has made them compatible with several flexible substrates, for example, plastic or paper substrates. This directly makes large-area, low cost, roll-to-roll, transparent and flexible systems promising [? ]. Prototypes of flexible displays, digital circuits and amplifiers have been reported in recent years demonstrating the possibility of realising modern electronics in a different and flexible manner [? ? ? ? ? ? ]. Several examples are shown in Fig. 1.3. Another promising approach for flexible electronics lies in healthcare. Lower power consumptions is a critical requirement in health care monitoring systems especially when 4 Introduction Fig. 1.3 Flexible circuits and displays with TFTs (a) stress sensor array with built-in amplifiers (b) Σ−∆ ADC (c) op-amp (d-f) flexible and transparent displays. Figures are captured from [? ? ? ? ? ? ]. continuous monitoring is needed in daily life, and in implantable applications where long lifetime is of concern. As TFTs are generally working at a lower current level compared with silicon MOSFET technology, they are naturally more suited for low power applications. Recent research has reported that ultra-low power working (<1nW) is possible [? ] when biasing TFTs in the subthreshold region, which could potentially lead to battery-less systems. It is worth mentioning that although many prototype applications have been demonstrated, building a full system on a flexible substrate is still very challenging as in most applications, rigid crystalline silicon-based integrated circuits or components are still needed in biasing and operating the TFT circuits, which are not currently replaceable with TFT circuits. 1.3 Research Goals To enable various functionalities efficiently, TFT circuit designs and simulations are indis- pensable. However, the simulation platforms for newly emerging technologies are not readily available. For a-Si TFTs, the well-known RPI model [? ] has become a commercial standard supported by many simulation platforms including SmartSpice, HSPICE, Spectre, etc where designers can easily simulate various circuits with adequate accuracy. However, as a newly emerging technology, no computer-aided design (CAD) tool for metal-oxide TFTs has yet become a standard despite the improved performances compared with a-Si. In order to accurately simulate circuits and design various functions with metal oxide TFTs, we need to first investigate both the static and dynamic models in a simulation en- 1.4 Thesis Organization 5 vironment. Subsequently, building blocks at the bottom level should be engineered to connect TFT designs with the well-researched silicon complementary metal-oxide-semiconductor (CMOS) circuit and systems. Therefore the main goal of the dissertation is as follows: 1. Investigate compact modelling for circuit simulations. 2. Investigate linear models for design simplicity, and develop an accurate small signal model specifically for TFTs. 3. Design building blocks for bottom level analogue circuits in order to build a bridge between the TFT circuit design with the well studied CMOS designs. 1.4 Thesis Organization The dissertation includes 5 chapters and 2 appendices. This first chapter introduces the background and the motivation of the research and explains the main goal of this research. Chapter 2 mainly reviews the CAMCAS model developed in Cambridge where a lot of work has been completed from previous efforts. Chapter 3 investigates the widely used CMOS small signal model and the inadequacy of its application in TFTs. A new model is developed taking into consideration of non-idealities in TFT devices. In Chapter 4, several building blocks for analogue circuits are designed and simulated with the CAMCAS model. The design of a supply independent op-amp is presented. Chapter 5 concludes this dissertation and discusses possible extensions of this research. The two appendices discuss the device circuit interaction (DCI) and subthreshold circuit possibilities of TFT devices. In particular Appendix A considers DCI that happens in TFT circuit designs and available approaches are reviewed. Thermal and geometric variations are measured and the effects on simple analogue circuit performance are analysed. Based on the knowledge developed in the above chapters, Appendix B evaluates the analogue performance and the design difficulties in the newly proposed way of biasing Schottky-barrier TFTs in subthreshold region, which extends the subthreshold modelling of Chapter 2. Chapter 2 TFT Compact Modelling 2.1 Computer Aided Design The growing maturity of thin film transistor (TFT) technology coupled with newly emerging materials and processes are enabling integration of circuits and systems for a new family of applications ranging from biosensing systems to areas augmenting displays and imaging [? ? ? ? ]. The design of systems places great demand for fast computer aided design (CAD) tools to accurately and reliably predict system behaviour. CAD always requires accuracy, high speed and reliable convergence of device models. In order to achieve high accuracy in all working regions without use of many fitting parameters, a good understanding of device physics is needed. However, physical models are not necessarily sufficient for good CAD in the sense that the equations could be complex, which can lead to slow execution or convergence problems. Therefore, simplification of the model equations without sacrificing accuracy is one of the major requirements for a good CAD model. In addition, the model should be adaptive to a wide range of technological/process parameters through simple extraction procedures. During the past 30 years, a great amount of effort has been made in the TFT modelling area for various semiconductor technologies, including amorphous silicon [? ? ? ? ? ? ? ], polysilicon [? ], organics [? ? ? ? ? ] and metal oxides [? ? ? ? ? ? ? ]. For a-Si TFTs, several SPICE models have been developed. Specifically, the model developed by Rensselaer Polytechnic Institute, also known as the RPI model [? ], has become a commercial standard supported by many simulation platforms including SmartSpice, HSPICE, Spectre, etc. However, no CAD tool for organic and metal-oxide TFTs has yet become a standard due to the rapid evolution of materials and device structures. In this chapter, we will review the major contributions to the TFT modelling area and compare the differences between Verilog-AMS and SPICE from the standpoint of compact 8 TFT Compact Modelling device modelling. Finally, we present the implementation of the Cambridge TFT compact model [? ] for oxide TFTs in Verilog-AMS for Spectre simulation. The major goal of this review is to give the reader guidelines for implementing device models for the CAD of circuits and systems. 2.1.1 Physical and Empirical Modelling Physical and empirical modelling approaches are generally aimed at providing an accurate model for a given transistor or transistor family, which covers all working regions of the device. While the major consideration is accuracy, the developed model is not necessarily physical. In this sense, it is theoretically possible to use polynomial approximation or other empirical ways to fit any kind of transistor behaviour with acceptable accuracy. However, without a proper understanding of the underlying device physics, the resulting number of fitting parameters could be large and difficult to extract from measurements since the model should cover a wide range of different bias conditions and transistor scales. Therefore, the transistor models used in a simulator are often a combination of terms and coefficients that are physically and empirically based. Recent efforts have primarily focused on physically-based approaches in an attempt to develop simple and accurate models. The most well-studied material for TFTs is amorphous silicon (a-Si:H). Several models have been developed based on different distributions of deep states and tail states of the semiconductor to describe static and dynamic behaviour for the above and subthreshold regions [? ? ? ]. Other properties including trap related VT shift [? ] and the off/leakage currents [? ]. The RPI model [? ] captures most of the device properties leading to satisfactory simulation results and thus has been widely used. For organic TFTs, researchers have developed models based on multiple trapping and release (i.e. trap-limited conduction) [? ] and variable range hopping (VRH) [? ? ? ]. However, due to the diverse use of materials in the TFT, the underlying physics differs, and therefore a trend has emerged to develop a model that is unified but less physical [? ]. As for metal oxide TFTs, modelling efforts are still in their infancy, although there are growing interests in the area for implementation of circuits and systems. In what follows, early efforts in compact models for CAD of oxide TFTs are reviewed. 2.2 Cambridge’s TFT Models Compared with a-Si TFTs, the oxide system has unique properties, which need to be captured. For example, localised traps or band tail states in oxides do not exist to the same extent as 2.2 Cambridge’s TFT Models 9 Fig. 2.1 Carrier transport combining percolation with trap-limited conduction (TLC) for oxide semiconductor TFTs (Here, DB and WB denote spatial distance and width of potential barriers, respectively) a-Si [? ]. Their tail state density is much lower and hence trap-limited conduction is generally insignificant [? ? ]. In addition, more complex systems such as amorphous indium gallium zinc oxide (a-IGZO) can have compositional disorder due to the random distribution of metal constituents [? ? ]. This gives rise to potential barriers above the conduction band minima (Em), suggesting the presence of percolation conduction [? ? ? ]. In the following, compact models for the terminal current-voltage behaviour are presented taking into account the different transport mechanisms in the device for the above- and sub- threshold regions of TFT operation. The former is based on a mobility model that combines trap-limited conduction (TLC) with percolation conduction. The latter takes into account diffusion and drift current components [? ]. A unified model is then presented that covers both regions based on a single expression that uses a reference turn-on voltage Von rather than VT [? ]. Good agreement with measured terminal characteristics is obtained over the entire range of gate-source voltage (VGS) > Von for the test TFTs with an a-IGZO channel. 2.2.1 Above-Threshold Model As illustrated in Fig. 2.1, oxide TFTs have potential barriers above Em due to compositional disorder, suggesting percolation conduction when electrons are released into the conduction band. Moreover, there are localised tail states within the gap states, implying trap-limited conduction. In particular, oxide semiconductors can have a shallow slope of the tail states (kTt) ∼ 20meV , smaller than the thermal energy (kT ) at 300K, leading to different mobility behaviour. This suggests that the field effect mobility (µFE) model needs to be modelled based on TLC and percolation conduction, although the former is not as significant as compared to a-Si. 10 TFT Compact Modelling 1. Trap-limited conduction (TLC) Effect of trap-limited conduction (TLC) can be considered as µFE being proportional to ratio (γT LC) of free carrier density (n f ree) and trapped carrier density (ntrap), yield- ing γT LC = n f ree/(n f ree + ntrap). Here, n f ree = NCexp[(EF −Em)/kT ], where NC is effective density of states in conduction band and kT the thermal energy. And the expression for ntrap is approximated with kTt < kT using exponential distribution of tail states. This yields ntrap = NtckTtexp[(EF −Em)/kT ], where Ntc is the density of trap states at conduction band edge. Now, we have γT LC as just a constant, γT LC ≡ n f ree (n f ree+ntail) ≈ NC NC +NtckTt (2.1) 2. Percolation Conduction Percolation conduction associated with potential barriers above Em can be consid- ered as mobility scaled from band mobility (µ0), assuming Gaussian random dis- tribution of potential barriers with the mean (φB0) and the variance (σB0). This yields µ∗0 = µ0exp[−qφB0/kT +(qσB0)2/(kT )2]. Here, φB0 can be reduced by ∆φB0 due to thermally released electrons, depending on Fermi level change (∆EF ), as described in Fig. 2.1. The thermally reduced barrier height can be expressed as φB0exp(−γB∆EF/kT ), where γB ≡ (DB−WB)/DB, which can be approximated as φB0(1− γB∆EF/kT ) when γB∆EF/kT << 1 by Taylor expansion. Thus, ∆φB0 is de- fined as γB∆EF/kTφB0. These conditions yield the percolation mobility (µPer) as follows, µPer ≡ µ0exp ( −q(φB0−∆φB) kT + (qσB0)2 2(kT )2 ) = µ∗0 exp ( γB∆EF kT φB0 ) (2.2) where µ∗0 = µ0exp[−qφB0/kT + (qσB0)2/2(kT )2] considered as an effective band mobility. 3. Combined Mobility Model with VGS dependence In Eq. (2.2), the ∆EF is controlled by gate voltage (VGS). The relationship between them can be derived by solving Poisson’s equation, yielding ∆EF = 2(kT/q)ln[Cox(VGS− VT )/Qre f ], where Cox is gate-insulator capacitance, VT threshold voltage, and Qre f ≡ {2εSNCkTexp[(EF0−Em)/kT ]}0.5. Combining TLC with percolation from Eqs. (2.1) to (2.2), we now have the VGS dependent mobility relation, 2.2 Cambridge’s TFT Models 11 µFE ≡ µPerγT LC = µ∗0 ( NC NC +NtckTt )( Cox Qre f )αp (VGS−VT )αp (2.3) As can be seen, Eq. (2.3) follows a power law. Here, αp ≡ 2qφB0γB/kT , related to percolation. In Eq. (2.3), TLC affects the constant term, while the exponent is determined by percolation. 4. Current-Voltage Relation With Eq. (2.3) and the definition of drift current: IDS = µFECox(VGS−VT−Vch)dVch/dx, current-voltage relation IDS(VGS) can be derived with the integral ranges: x=0 to L and channel potential (Vch) = 0 to VDS, IDS ≡ µ∗0 ( NC NC +NtckTt ) W L′ Cαp+1ox Qαpre f (VGS−VT )αp+1V ′DS (2.4) In Eq. (2.4), L′ is defined as L−∆L, where ∆L is effective channel length reduction, and effective drain voltage V ′DS =VDS−2RCIDS, where RC is contact resistance. For saturation region expression, Eq. (2.4) can be reformed with a saturation parameter (βsat), replacing V ′DS by βsat(VGS−VT ). Fig. 2.2a shows a comparison between mea- sured and modelled transfer characteristics (IDS vs. VGS) at VDS = 0.1V , providing a good agreement. The measured saturation characteristics at VDS = 20V is also well matched with the modelled results, as seen in Fig. 2.2b. To check the model validity for temperature dependency, we measured and simulated the drain current as a function of gate bias for different temperatures, e.g. 100K, 200K, and 300K, respectively. As seen in Fig. 2.3a, there is good agreement between the measurements and the model, Eq. (2.4). For purposes of validation, the modelled values of exponent (αp) in the power-law, Eq. (2.4), are compared with the extracted values from the best fit since it has a unique signature of percolation conduction, i.e. a = αp + 1 ≡ 2qφB0 · γB/kT + 1. As seen in Fig. 2.3b, the proposed percolation model for φB0 · γB = 2.5meV shows better agreement compared to the conventional model (e.g. trap-limited conduction model with a = 2kTt/kT −1 for kTt = 30meV ). 2.2.2 Subthreshold Model The subthreshold current in the limit of low VGS shows a linear dependence on VGS in a semi-log plot, as illustrated in Fig. 2.6, suggesting diffusion current, as follows, 12 TFT Compact Modelling (a) linear region with a small drain voltage (VDS) of 0.1V (b) saturation region with a high VDS of 20V Fig. 2.2 Comparison between measured and modelled transfer characteristics (IDS vs. VGS) for the above-threshold region. For each region, a good agreement between measurement and modelling is achieved. 2.2 Cambridge’s TFT Models 13 (a) Measured and modeled IDS-VGS for above-threshold region at VDS = 0.1V for different temperatures (100K, 200K, 300K, respectively) (b) Retrieved exponent (a) vs. temperature Fig. 2.3 Comparison between measured and modelled transfer characteristics at different tem- peratures. Here, the proposed model shows a better agreement compared to the conventional model (e.g. trap-limited conduction model). 14 TFT Compact Modelling Fig. 2.4 Density of states profile with (a) Fermi level within interface states at low gate voltage, and (b) the case of deep states dominant with increasing gate voltage corresponding Fermi level now within deep states. IDi f f ≈ µ0 kTq W L′ Q f iexp ( q kT ( VGS−VFB 1+q2Dit/Cox ))( 1− exp ( −qV ′ DS kT )) (2.5) where µ0 is the band mobility for electrons. Note that the subthreshold slope (SS) can be derived from Eq. 2.5 with the definition dVGS/dlogIDS, S≡ ln10kT q ( 1+ q2Dit Cox ) (2.6) As described in Fig. 2.4, interface states are occupied first, followed by filling deep states located in the bulk at higher VGS. So, as the EF moves to the location of deep states for increasing VGS, the drain current can be defined as a drift current (IDri f t), IDri f t ≈ µ0WL′ Cαd+1ox Qαdd (VGS−VFB)αd+1V ′DS (2.7) where Qd is a reference charge density associated with deep states, αd is power-law exponent defined as 2(Td/T −1), and Td is the characteristic temperature of deep states. We now have the current-voltage relations for both diffusion and drift components as given by Eqs. (2.5) & (2.7). These equations can be combined as a total drain current (Isub) in the subthreshold region using a harmonic average, Isub ≡ (I−mDi f f + I−mDri f t)−1/m (2.8) 2.2 Cambridge’s TFT Models 15 (a) linear region with VDS = 0.1V (b) saturation region with VDS = 20V Fig. 2.5 Measured and modelled subthreshold current (Isub) as a function of VGS 16 TFT Compact Modelling Table 2.1 Extracted Subthreshold Parameters at T=300K Parameters Value SS 0.19V/dec Dit 1.64×1011cm−2eV−1 Io f f 2×10−14A Q f i 7.05×10−14C/cm2 αd 2.26 Td 639K Qd 7.8×10−8C/cm2 The examined TFT to verify this model, which is the same as that used in the previous sections, has the following geometrical and physical parameters (extracted in the previous sections): W = 100µm, L = 100µm, channel thickness tS = 40nm, channel permittivity εS = 11.5ε0 (where ε0 is vacuum permittivity), VFB ∼ 0.6V , VT ∼ 4V , Cox = 11.5nF/cm2, µ0 = 15cm2/V · s, and NC = 5×1018cm−3 at 300K. 2RC ≡ RSD = 9637Ω for W = 100µm (equivalent to RSDW = 96.37Ω · cm) and ∆L = −3.5m (L′ = L−∆L = L+ 3.5µm). Other extracted model parameters are summarized in Table 2.1. Fig. 2.5 shows the measured subthreshold characteristics for a different VDS, providing a good agreement with each other. To validate the subthreshold model in terms of the diffusion component in the examined transistors, we measured the drain current for small VDS=0.01, 0.1, and 1V. Comparing to the model, we get good agreement as seen in Fig. 2.6. In particular, as shown in the inset of Fig. 2.6, the measured IDS vs. VDS at the diffusion dominant region of VGS (e.g. VGS = 0.5V) is compared with the Eq. (2.5). Here, the dependence of the drain current on VDS is found to follow the law of (1− exp(−qVDS/kT )) of Eq. (2.5) which is a signature of the presence of the diffusion current. Regarding continuity and symmetry of the compact model where above- and sub- threshold models are combined, the model was subject to the Gummel symmetry test (GST) [? ? ]. As seen in the inset of Fig. 2.7(a), VX represents a symmetrical voltage applied on source and drain sides, respectively. Fig. 2.7 demonstrates perfect continuity and symmetry as a function of VX even for the 4th derivative of IDS with respect to VX , suggesting it has successfully passed the GST. For this, we employed smoothness and continuity functions for the effective drain voltage and threshold voltage terms [? ]. 2.2 Cambridge’s TFT Models 17 Fig. 2.6 Measured and modeled IDS vs. VGS at subthreshold region for a different VDS. Inset: Measured and modeled IDS vs. VDS for VGS = 0.5V . Here, the equation is simplified from Eq. (2.5) introducing the pre-constant I0 which is around 10pA at VGS = 0.5V. 2.2.3 Off Region Model The off current can defined according to [? ] at VGS =VFB, Io f f ≈ µ0 kTq W L′ Q f i ( 1− exp ( −qV ′ DS kT )) (2.9) 2.2.4 Unified Model As seen in the previous sections 2.2.1 & 2.2.2, the current-voltage relation needs to be derived separately to describe the subthreshold and above-threshold characteristics. Here, we need separate expressions for the subthreshold and above-threshold regions, implying two different equation systems to describe total current (see Fig. 2.8). Moreover, in this case, threshold voltage (VT ) is not immediately apparent from the I-V plot and needs to be extracted from above-threshold region of the characteristic as a fitting parameter. However, the extracted value of VT can be quite different depending on extraction method and the I-V data range chosen for the fit. In contrast, turn-on-voltage (Von) is the gate voltage (VGS) at which drain current (IDS) starts increasing rapidly, thus it is easy to identify Von on a semi-log plot of IDS vs. VGS. 18 TFT Compact Modelling Fig. 2.7 (a) Calculated IDS vs. VX of the combined above- and sub-threshold model for different VG (4, 6, 8V). (b) First, (c) second, (d) third, and (e) fourth derivatives of IDS with respect to VX . The inset of (a): test circuit configuration of the GST. 2.2 Cambridge’s TFT Models 19 Fig. 2.8 (a) Schematic IDS vs. VGS curves (gray circles) with the conventional power- law models for subthreshold (Sub-T, solid line) and above-threshold (Above-T, dot line) characteristics. In this case, we need two models for these two different operational regions. Here, Von and VT are on-voltage and threshold voltage, respectively. (b) Schematic IDS vs. VGS curves (gray circles) with only one model: unified model which can cover subthreshold as well as above-threshold regions at the same time. We can unify this to cover both subthreshold and above-threshold characteristics. The proposed unified model is a single expression with a reference voltage level Von rather than VT , providing good agreement with measured terminal characteristics over the entire range of VGS >Von for the test TFTs with an amorphous InGaZnO (a-IGZO) channel, which is the same TFT used in the previous sections. The derived equation for this model is as follows, IDS = G0 W L′ exp(κ(VGS−Von)α)V ′DS+ Io f f (2.10) where G0 = µ0(εSkT NC)1/2, κ = ζ/2kT , and ζ & α are related to trap states. To extract the values of κ and α , we can rewrite Eq. (2.10) as, U ≡ I ′ DS/V ′ DS d(I′DS/V ′ DS)/dVGS = 1 ακ (VGS−Von)1−α (2.11) Here, I′DS ≡ IDS− Io f f . Therefore, 1−α and 1/(ακ) can be extracted from the plot of lnU vs. ln(VGS−Von) as the slope and the intercept, respectively (as illustrated in Fig. 2.9a). The extracted model parameters are summarized in Table 2.2. Fig. 2.9b shows modelled results for different L (=25, 50, and 100µm), providing good agreements with the measured characteristics. Also, they exhibit small average errors < 5% over a wide range of VGS from 5 to 20V. Interestingly, the proposed model using only one equation covers both the sub- and above-threshold regions at the same time. This is mainly 20 TFT Compact Modelling Table 2.2 Extracted Parameters for Unified Model at T=300K Parameters Value G0 2.34×10−5ohm−1 κ −10.812V−α α -0.675 due to a combination of the exponential function and power-law. Additional advantage of this unified model is that it just needs a few model parameters to be implemented in model code description, e.g. Veriog-A, thus providing a higher-speed simulation. 2.3 Computer Interpretation of Device Model After having established an accurate analytical model for the TFT, computer implementation for circuit simulation is not always easy. Several considerations including speed, convergence and even the way of translating different form of equations into computer language should be taken into consideration in this step. For a-Si TFTs, several SPICE models have been developed. Specifically the model developed by Rensselaer Polytechnic Institute, also known as the RPI model [? ], has become a commercialized standard supported by many simulation platforms including SmartSpice, HSPICE and Spectre, etc. However, CAD of organic and metal-oxide TFTs is yet to become a standard due to the rapid and ongoing evolution in materials and device structures. In fact, and as mentioned previously, there is a quest for unifying the TFT model to meet simulation requirements of TFTs with ever changing structures and materials. It is important to understand the benefits and drawbacks of SPICE and Verilog-A, while developing a device model to be used in further circuit simulation and system design. 2.3.1 SPICE SPICE is an open source analogue circuit simulation software firstly developed at the University of California, Berkeley [? ]. The first two versions of this simulator was written in Fortran. SPICE2 included several elements widely used in circuit simulation at that time, including the diode, MOSFET, JFET and BJT, etc. In 1989, a version of the simulator SPICE3 was re-written in C. This version includes the well-known BSIM model [? ]. Many commercial circuit simulators today are based on the aforementioned software including PSPICE, HSPICE, SmartSpice, Spectre, etc. The fast simulation speed, high accuracy and ease of use make the SPICE simulator extremely popular in the area of 2.3 Computer Interpretation of Device Model 21 (a) Plots of lnU vs. ln(VGS−Von). Here, linear fitting is applied over the entire range of ln(VGS−Von) (b) Modeled results for different L in comparison with the measured curves using a unified model, normalized error is illustrated in the inset. Fig. 2.9 Parameter extraction for unified model & measured and modelled transfer curve with percentage error 22 TFT Compact Modelling circuit design. However, SPICE is written in a low level computer language (Fortran or C) and so are the models used in it. Therefore, researchers should not only have a good understanding of device physics but also need enough knowledge of the algorithms and limitations of the simulator. The procedure for developing a stand-alone SPICE model (not by combining existing elements) always takes time and resources. Due to the complexity of model development in SPICE, currently only well developed and widely used devices are having SPICE models. It is worth noting that modern circuit simulators use a heavily modified original SPICE model but they separate as much as possible the simulation algorithm from the device model. For example, the Spectre simulator provides a Spectre Compiled Model Interface (CMI) for Spectre developers and others to use C code directly for device modelling. 2.3.2 Verilog The Verilog Hardware Description Language (Verilog HDL) was designed in 1984 to meet the design demands of VLSI engineers working in the higher abstractive level to increase the number of transistors in simulating digital circuits. Subsequently adopted by Cadence Design Systems and Open Verilog International, VHDL has become an IEEE standard in 1995, i.e. IEEE Std 1364-1995 [? ]. Since the original Verilog language did not support analogue systems, the extensive use of HDL languages led to a growing demand for high-level behavioural model for systems and components for use in analogue and mixed signal system design. Thus, Verilog-AMS standard was then developed by Accellera to include Verilog-A, a previously standalone behavioural language used for analogue modelling in Spectre, and extending both the digital part of the IEEE Std 1364 and the analogue part of Verilog-A [? ]. In the early days of Verilog-A and Verilog-AMS, the behaviour model was limited in terms of accuracy for analogue devices. The language was first developed for design of systems and thus the behavioural language was at that time not intended for complex physical equations at the device level. Fig. 2.10 illustrates the mixed signal simulation environment in early days. In addition, device modelling often led to long execution times compared with SPICE models. Even so, the capability of behavioural modelling started to attract the attention of device modellers struggling with low level implementation. In 2002, the open-source tool supporting automatic conversion [? ] of Verilog-AMS models to C codes was introduced. Since then, several proposals have been made to use Verilog-AMS to efficiently develop compact device models [? ? ] to integrate with SPICE like simulators. The report in 2010 [? ] showed comparable performance of their Verilog- AMS device model against SPICE. The language has now become a major tool for device 2.3 Computer Interpretation of Device Model 23 Fig. 2.10 Verilog-A behaviour model in early mixed signal simulation environment Fig. 2.11 Verilog-A(MS) extension for compact modelling with SPICE like simulator inte- gration 24 TFT Compact Modelling Table 2.3 Comparison between different model languages SPICE Spectre CMI Verilog-AMS Language C/Fortran C Verilog Language level Low Low High Easy to use No No Yes Efficiency & Speed High High Low Complexity High High Low Simulator information needed for developers Simulation algorithm & limitations Simulator limitation on equations None Users SPICE developers CAD engineers, Cadence programmers End users model developers. Fig. 2.11 illustrates the relation between different Verilog languages and the simulation flow in two types of simulators. 2.3.3 Short Summary of Comparison Table 2.3 summarizes the comparison of different model languages. The SPICE model development based on a low level language is only used by model developers of components, which are not likely to change (due to heavy usage or technology maturity) in view of resource requirements. In the meantime, the Verilog approach is ideally suited to meet the modelling demands of ever-changing TFT or other device structures without requiring a mature physical understanding of the devices or underlying material system. The direct implementation of physical equations in Verilog-AMS will dramatically reduce the time between development of new devices and subsequent circuit simulation. Despite all the benefits of Verilog-AMS, the model is still slower and less efficient than a well-developed SPICE model, due to the fact that the Verilog-AMS model should be converted to a lower level language form. The converted form although automatically optimized by the synthesizer (e.g. ADMS [? ]) cannot be as efficient as a SPICE model because the latter is optimized directly in a lower level language. Therefore, it is recommended to use SPICE for circuit simulation whenever it could offer sufficient accuracy. On the other hand, Verilog-AMS is suited for devices that are new or under-development. 2.4 CAMCAS Model in Simulation Environment 25 Fig. 2.12 Circuit design and simulation flow using CAMCAS model in Cadence Design Environment 2.4 CAMCAS Model in Simulation Environment As discussed before, we implement the Cambridge’s TFT compact model (i.e. CAMCAS model) directly into Verilog-A language for use in a Cadence Design Environment so as to focus on developing a more accurate model for circuit simulations. The simulation flow in the Cadence Design Environment is shown as Fig. 2.12. The user extracts processing parameters based on the measurement results of their own devices and incorporates them into the CAMCAS Verilog-AMS model. In addition, the user can change the form of equations for better fitting to measurement results. Thus circuit designers can use the usual route of circuit design by simulating their TFT circuits before fabrication analogous to using the SPICE model in Cadence Spectre. Fig. 2.13 shows an example of an IGZO TFT circuit simulated using the CAMCAS model in Cadence Spectre. The simulation results show a voltage gain of 12dB in both transient and AC simulation as shown in Fig. 2.14a and 2.14b, respectively. 2.5 Summary and Discussion Physically-based compact models play an important role in computer-aided circuit design and simulation. This chapter reviews the major contributions in TFT analytical modelling and compares the pros and cons of SPICE and Verilog-AMS from the standpoint of device 26 TFT Compact Modelling Fig. 2.13 A voltage amplifier used in simulation with CAMCAS model model development. Brief guidelines are provided to decide on choice of modelling language for implementation of models and to understand how models work within the simulator. Regarding the device models, oxide TFTs are considered and two different approaches are reviewed. While considering physical and material properties of oxide semiconductors, both above- and subthreshold models have been presented. In particular, the above-threshold model uniquely addresses carrier transport properties in oxide TFTs, such as percolation conduction along with trap-limited conduction. Thus it is a fully-physical model. Also, using a harmonic average technique, diffusion and drift current components have been seamlessly combined to describe subthreshold current behaviour, providing good agreement with measurements. Using a different, more empirical-based approach, a unified model is discussed that covers both sub- and above-threshold regions with a single expression. This model provides higher speed circuit simulation since it has only a few number of parameters. The results presented here are essential for oxide circuit design and simulation, providing designers with a preference for fully physical but slower simulations on the one hand, and higher speed but less model complexity on the other. The CAMCAS model described above is implemented in Verilog-AMS, and the validity of the approach is demonstrated by way of a voltage amplifier circuit simulation. 2.5 Summary and Discussion 27 (a) Transient analysis for small signal amplitude at input and output (b) AC analysis for gain and phase Fig. 2.14 Simulation results of the voltage amplifier Chapter 3 Small Signal Modelling 3.1 General Small Models Small-signal model is used for linear approximation of many none linear electronic devices including diodes, MOSFET and BJT, etc. when the signal under processing is reasonably small compared with the DC bias of devices. The use of small-signal model significantly simplifies the design procedure of many analogue circuits, especially amplifiers. Different small-signal models exist for different electronic devices, for example MOSFET and BJT are having different small-signal models due to their different structure and working principles. However, researchers in the area of TFT circuit design are using the same small-signal model as MOSFET. In spite of the similarity of their structures, TFTs are having a much bigger contact resistance between different materials, more significant parasitic capacitance and a severe instability problem. Therefore, it is helpful to reevaluate the small-signal model specifically for TFTs. 3.1.1 Midband Small Signal Model First, we review how a midband small signal model was developed. As linear systems are to be designed out of none linear devices, the straightforward solution is to use a small part of the none linear behaviour (I-V curves) to approximate a linear response (small signal model). Therefore, under the assumption that the signal is sufficiently small, the linear approximation would be equal to the derivative of the none linear device at the bias point. For a general 3 terminal device with terminal G, S and D (terminal names can be arbitrarily chosen, here 30 Small Signal Modelling they are chosen in consistency with TFTs), The I-V formula can be written as:{ ID = f1(VGS,VDS) IG = f2(VGS,VDS) (3.1) Here, VGS and VDS are the voltage drops between G-S terminals and D-S terminals, respec- tively. As VGD can be calculated from VGS−VDS and IS =−ID− IG, Eq. (3.1) is enough to represent the I-V of an arbitrary 3 terminal device. For small changes of VGS and VDS, the current differences would approximately follow the derivative of the ID and IG functions in Eq. (3.1). dID = ∂ f1 ∂VGS dVGS+ ∂ f1 ∂VDS dVDS dIG = ∂ f2 ∂VGS dVGS+ ∂ f2 ∂VDS dVDS (3.2) For sufficiently small signals and sufficiently good linearities at the bias point, the partial derivatives can be treated as a constant. The small signal relations should follow: id = gm1vgs+ vds ro1 ig = vgs ro2 +gm2vds (3.3) where i and v in small cases represent the small change in current and voltage of the corresponding terminals, respectively. The coefficient of vds for id or vgs for ig can be represented as passive components (a resistor connection), while the coefficient of vgs for id or vds for ig can only be represented by active components, a voltage controlled current source (VCCS). Therefore, resistance is used to represent passive components (i.e. ro1 and ro2) and transconductance is used to represent active components (i.e. gm1 and gm2). Now, from Eq. (3.3) we have the linear approximation of the device at a fixed bias point of VGS and VDS (Note that the coefficients in Eq. (3.3) can be a function of VGS and VDS, but not a function of vgs and vds under the small signal assumption.) The corresponding small signal model of the equation can be drawn as Fig. 3.1. This is a general form of a hybrid-pi model which is consistent with both BJT and MOSFET. In the case of most transistors, VDS does not affect IG. Thus, gm2 = 0. And for the case of MOSFET, the gate resistance is huge and normally treated as open-circuit. Therefore, the corresponding components can be removed. The analysis reviewed above is a general approach for all three terminal devices. Thus it should also be valid for TFTs at midband frequencies. However, as the parameters are derived 3.1 General Small Models 31 Fig. 3.1 Small signal model for a general 3 terminal device from the derivative of I-V curves, a more physical representation of those parameters can be complicated and some of the parameters should be represented as separate components. 3.1.2 High-Frequency Small Signal Model An accurate calculation of the circuit response at midband frequency is not enough for circuit designers as the bandwidth and phase margin should also be predicted when designing circuits. This is especially crucial for amplifiers and active filters to prevent self-oscillation. While midband small signal models can be developed through a general way from I-V curves, the high-frequency small signal model can only be developed by considering I-V, C-V and physical meanings of different components. This is why the high-frequency small signal models are very different for CMOS and BJT despite the similarities on their midband models. The comparison of CMOS and BJT models is shown in Fig. 3.2 [? ]. The CMOS high frequency model (Fig. 3.2a) is more consistent with the midband model that the parasitic capacitors are added to the midband model while keeping the original connections. In comparison, the BJT model (Fig. 3.2b) is less so as node B′ is added between rx and rπ (which are the same resistance in the midband model). The separation of rx is due to the physical representation of a base spreading resistance which is the resistance of the connection to the base inside the device. The contribution of this resistor becomes more significant at high frequency because of the reduced impedance of the parasitic capacitors, namely Cµ and Cπ . 32 Small Signal Modelling (a) CMOS model (b) BJT model Fig. 3.2 High frequency small signal models for CMOS and BJT The review of CMOS and BJT high-frequency small signal models suggests that physical structure should be considered while developing high-frequency models as the connections of parasitic capacitances may vary and affect the overall performance of the model. Therefore, it is imperative to evaluate the TFT structure to include physical resistance, capacitance and their connections correctly. 3.2 TFT Small Signal Model As discussed in the introduction section, an accurate TFT small signal model needs to consider the physical representations and modelling not only for the sake of a simpler 3.2 TFT Small Signal Model 33 expression of each component in the midband model but also for an accurate modelling in the high-frequency model. In this section, we will discuss the concerns at the stage of model developing and report the developed models. 3.2.1 Midband TFT Model As discussed in section 3.1, the midband model can be developed using the general approach. Therefore, the small signal behaviour can be represented by a VCCS and an output resistor in parallel. The major concern of this approach is the expression of the two components, gm and ro. As the two components come from the derivative of IDS−VGS and IDS−VDS curves, the static characteristic of TFTs should be considered first before deriving the expressions of gm and ro. The major concern here comes from the big contact resistance (RC) [? ? ? ? ? ? ] and the threshold voltage shift (VT shift) [? ? ? ? ? ? ] in TFTs which include TFTs from different material families such as amorphous silicon, organic and metal oxide semiconductors. Although TFTs are having a similar working principle with MOSFET, the contact resistance (RC) is so dominant that even the static model should consider RC as independent components. Most of the static models are developed based on a structure shown in Fig. 3.3 where the I-V characteristics are modelled for the internal transistor and the RCs (RS and RD here) are accounted for separately. In this section, we will consider the effect of contact resistance and VT shift in the midband TFT model. The effect of contact resistance As the overall I-V expression is complicated when considering RS, RD and the ideal TFT as a whole, it is more convenient to derive the small signal model of the ideal TFT and then consider the circuit as a linear network. Therefore, by using the I-V expression we can easily derive a small signal model shown in Fig. 3.4a. As reviewed in Chapter 2, the general I-V expression for a TFT can be written as: IDS = 1 αp+2 K W L′ (VGS′−VT )αp+2 (3.4) where K = (αp + 2)µ∗0 ( NC NC+NtckTt )C αp+1 ox Q αp re f βsat . The transconductance gmi in in Fig. 3.4a can then be calculated directly as: gmi = K W L′ (VGS′−VT )αp+1 (3.5) 34 Small Signal Modelling Fig. 3.3 The inner structure based on which static TFT models are developed The roi in the figure, however, is harder to express in physical parameters. As the slope of the output characteristics are normally modeled using an empirical Early voltage (VA)1, roi is then calculated using: roi = VA IDS (3.6) With the above equations, we can easily get an accurate midband small signal model for the TFT of concern and design suitable gmi and roi by choosing W/L ratio, given that the TFT technology and processing parameters are already modeled with known RS, RD and VA. However, it is not so convenient in terms of measurements as the results will be overall I-V curves directly. As the derivatives of the overall I-V curves can capture the overall gm and ro 1TFTs do not have the same short channel effect as CMOS does. But the output characteristics have similar trend of having an Early voltage. Thus, it is usually modeled by an empirical Early voltage. 3.2 TFT Small Signal Model 35 (a) RC as separate component (b) Overall midband model Fig. 3.4 Midband small signal models in a equivalent midband model shown in Fig. 3.4b (the general approach). It is worth finding the relations between the two models shown in Fig. 3.4. As the circuits are only consists of linear components, in this case only VCCS and resistors, the circuit in Fig. 3.4a should be equivalent to the one in Fig. 3.4b according to 36 Small Signal Modelling Thévenin’s theorem [? ? ]. From Fig. 3.4a, we have: id = vs′s RS id = gmi(vgs− vs′s)+ vds− idRD− vs′s roi (3.7) Substituting vs′s and considering 2 RS = RD = RC, we get: id = ( gmiroi gmiroiRC +2RC + roi )vgs+( 1 gmiroiRC +2RC + roi )vds (3.8) Compared with Fig. 3.4b where we have: id = gmvgs+ vds ro (3.9) As both models should capture the same derivatives of the overall I-V curves (i.e. gm and ro), the relations should follow: gm = gmiroi gmiroiRC +2RC + roi ro = gmiroiRC +2RC + roi (3.10) Here, we can derive the overall gm and ro out of the theoretical prediction of our transistor models. While designing real circuit applications, this will be the equations to use to estimate the circuit behavior. By solving Eq. (3.10), we can also get the values of internal components from measurement results: gmi = gmro ro−gmroRC−2RC roi = ro−gmroRC−2RC (3.11) This equation is very useful when the values of internal components are to be extracted out of measurement results. It also provides a convenient way of checking the validity of the model, which will be discussed later. Eq. (3.10) and Eq. (3.11) not only show the relation between the static model param- eters and the overall parameters from measurements but also reveal some insights of the parameters: 2As the semiconductor layer and source/drain contacts are made of the same material and have identi- cal geometries and similar properties, the contact resistances are considered to be equal at both sides for convenience. 3.2 TFT Small Signal Model 37 1. The intrinsic gain gmro of both the overall performance and the ideal internal TFT are the same. As the overall gm is reduced due to the contact resistance (in comparison with gmi), the ro increases at the same rate making gmro = gmiroi. This also means that the contact properties of the semiconductor layer and the electrodes do not affect the intrinsic gain of the overall transistor. The measurement of gmro directly shows the maximum achievable single-stage gain of the semiconductor technology. 2. The term gmiRC can be used to evaluate how different the measurement results will be from the ideal TFT in the model. As gmiroiRC cannot be dominant over roi, otherwise the overall gm ≈ 1RC (which is impractical as there’s no dependence on bias), and RC << roi, the value of gmiRC should be around 0 1 (might be a bit higher than 1 but should be around the order of magnitude). The effect of VT shift Another property that might affect the midband model is the VT shift under bias condition. As the implementation of TFTs in AMOLED display proves that bias-induced VT shift must be compensated especially when the analogue properties are to be used (i.e. the output current needs to be actually controlled), the TFT small signal model should capture the behaviour to help design and evaluate TFT analogue circuits. The major concern of VT shift is that it would change the bias condition in effect, as in the static models, the VT terms always appear in VGS−VT . Therefore, considering VT shift, the expression of Eq. (3.5) can be rewritten as: gmi = K W L′ (VGS′−VT −∆VT )αp+1 = K W L′ (VGS′−VT )αp+1 ( 1− ∆VT VGS′−VT ) = K W L′ (VGS′−VT )αp+1−K W L′ (αp+1)(VGS′−VT )αp∆VT +O ( ∆VT VGS′−VT ) (3.12) Typically, ∆VT << VGS′ −VT , as the experiments show that the VT shift does not turn the transistor off (i.e. ∆VT ∼ VGS′ −VT ) at least not within a reasonable time period. Hence, the higher order terms can be omitted in the above Taylor expansion. The first order approximation of gmi can then be written as follows: gmi ≈ gmi0−K′∆VT (3.13) 38 Small Signal Modelling Fig. 3.5 Midband TFT model considering VT shift where gmi0 = K WL′ (VGS′ −VT )αp+1 is the initial value where there’s no VT shift, and K′ = K WL′ (αp+1)(VGS′−VT )αp is the coefficient of the gmi change per ∆VT change. Here, as αp is a constant value close to zero, K’ can normally be treated as a constant coefficient. This means that the VT shift effect can be represented by a separate component. The corresponding modification of small-signal model is shown in Fig. 3.5. 3.2.2 High-Frequency TFT Model Device structure and physical meanings of all components need to be considered in high- frequency small signal model, especially when capacitances come into the picture, as dis- cussed in section 3.1 of the BJT model. The connection of contact resistances in TFTs plays a similar role as the base spread resistance of BJT which is connected in series with a parasitic capacitance. Here, as illustrated in Fig. 3.6 the contact resistance at source side (RS) separates the overlap capacitance (COV S) and channel capacitance(Cch). Hence, RS might not be negligible nor included in overall components (i.e. gm and ro) especially at high frequencies. Combining the analysis of the midband model, the high-frequency small signal model for TFT can be drawn as Fig. 3.7a. In comparison, the CMOS model adapted to TFT parameters are shown in Fig. 3.7b. Here, COV S and Cch are considered to be the same capacitor between source and gate terminals and the contact resistances are included in the overall gm and ro. 3.2 TFT Small Signal Model 39 Fig. 3.6 Bottom-gate TFT structure considering passive components of the small signal equivalent circuit. COV S and COV D are the overlap capacitance at source and drain side; Cch is the channel capacitance; RS and RD are the contact resistances at source and drain side. (a) TFT model (b) CMOS model adapted to TFTs Fig. 3.7 High frequency small signal models 40 Small Signal Modelling In order to evaluate the difference between the developed model and the CMOS model, we compare the expressions of the short circuit current gain (Ai) and the cutoff frequency of the two different circuits. For CMOS model, Ai can be calculated as: Ai = sCOV D−gm s(COV D+COV S+Cch) (3.14) which contains one pole (at 0) and one zero (at gm/COV D). Assuming the zero is far from the point of cutoff frequency. The unit-gain cutoff frequency can be calculated as: fT = gm 2π(COV D+COV S+Cch) = gm 2πCtot (3.15) where Ctot is the sum of all capacitances of the TFT. In contrast, the current gain calculated for the TFT small signal model seen in Fig. 3.7a is Ai = −gmiroi+((2RC + roi+gmiRCroi)COV D+CchRC)s+COV DCchRC(RC + roi)s2 s(gmiCOV RCroi+(2RC + roi)(COV +Cch)+COV (RC + roi)RCCchs) (3.16) where COV =COV S+COV D is the total overlap capacitance. The equation shows two zeros and two poles for this model. Assuming the poles and zeros are far from the cutoff point (for linear assumption at the point in Bode plot), the cutoff frequency expression is derived as: fT = gmiroi 2π (gmiCOV RCroi+(2RC + roi)(COV +Cch)) = gm 2π(Ctot −gmRCCch) (3.17) The simplification of the above equation shows that the major difference between Figs. 3.7a and 3.7b comes from the term gmRCCch. As gm is a term already contains the RC, it is preferable to consider this term as a function of independent variables. Thus, by substituting the gm term, the expression of gmRC is: gmRC = gmiroiRC gmiroiRC +2RC + roi (3.18) As gmi, roi and RC are independent values depending on semiconductor and interface proper- ties with positive values, the whole term should be smaller than one. This indicates that the fraction Cch is of the total capacitance plays an important role in determining whether the two models become identical. In the case when COV is dominant over Cch or when the value of RC is very small or close to zero, the gmRCCch term can then be omitted and thus Eq. (3.17) becomes the same as Eq. (3.15). The CMOS model can be used despite the connection of RCs as in these cases the two high frequency models are equivalent. 3.3 Model Validation 41 In the case when the value of Cch is at least of the same order of magnitude as COV or higher, the gmRCCch term can reduce around 50% of the total capacitance (for the case when gmiRC ≈ 1) making the two models very different in terms of modelling of the cutoff frequency. This also indicates that the CMOS model can be inaccurate especially when the contact resistance and overlap capacitance are big and small, respectively. As the contact resistance is known to be big in TFTs (from a few kΩ to several 100kΩ) and the reduction of overlap capacitance is preferable at the processing side for the sake of making faster TFTs, the inaccuracy of CMOS model might exist in most TFTs and may become more significant as processing technique becoming mature (e.g. for self-aligned TFTs the overlap capacitance is close to zero). The other point to notice is that the assumptions used to derive Eq. (3.17) is supported by calculating the actual values of zeros and poles of the transfer function. As the extra zeros and poles are at way higher frequency compared with the cutoff frequency, they are not of concern as the transistor is dominated by capacitor effects and is not working properly at the frequency range. 3.3 Model Validation From the theoretical analysis of small signal modelling, we know that the TFT small signal model can be very different from CMOS model when the contact resistance and overlap capacitance are big and small, respectively. In this section, we will evaluate the amount of difference by measurement of IGZO TFTs’ cutoff frequency and the dominating s-parameters. 3.3.1 Measurement Setup In order to evaluate the small signal model, we need to measure the current gain of the TFT and compare the cutoff frequency to see the model difference quantitatively. The common way of measuring the cutoff frequency of a transistor is to use a network analyser to measure the s-parameters and convert them to h-parameters. The calculated h21 directly reflect the short-circuit current gain of the device. Here, we use Keysight E5061B network analyser (ENA) calibrated by CS-11 calibration substrate provided by GGB Industries, Inc. The standard measurement setup is shown in Fig. 3.8. The bias-T provides DC bias to the TFT under test and separates the small signal input and output from the DC circuitry. The TFT is then connected to the ENA ports for s-parameter measurement. The main reason of choosing the analyser is that it can provide low-frequency measure- ment down to 5Hz. In contrast, most of the network analysers on the market are targeted 42 Small Signal Modelling Fig. 3.8 Measurement setup for fT and s-parameters above 100kHz range, which is already a range close to TFT’s cutoff. In the next section, we will see that even with the network analyzer low-frequency measurement is still hard, as the TFTs are more resistive at low frequencies. The ENA in use can only provide DC bias at port 1. Therefore, the bias-T at port 1 is already included inside the ENA. We still need to find a bias-T for port 2. As low frequency is important for our measurement, we choose Picosecond 5546 bias-T for this purpose. The low 3dB frequency is 3.5kHz, the lowest we managed to find on the market. We choose CS-11 calibration substrate and model 10 high-frequency probe (from DC to 6GHz) from GGB Industries, Inc. for the purpose of calibration, as the parasitics on the wires and the probes, etc. should be calibrated out before s-parameter measurement. The DC bias conditions are chosen to be VGS = 8V VDS = 15V to make sure the TFT is working in the saturation region. It is to be noted that from the output characteristic (Fig. 3.9) the TFT saturates later than VGS−VT , which is the saturating point for CMOS. This is explained as the voltage drop on the contact resistance effectively reduced the VD′S′ on the internal TFT (Fig. 3.3). The device under test has following physical and geometrical parameters: ts = 50nm, VT = 1.6V , Cox = 30nF/cm2, µ∗0 = 8.6cm 2/V · s. The device with W/L ratio of 100µm/10µm is used for model validation and 50µm/10µm is used for stress measurement. 3.3.2 s-parameters and fT Measurement As fT is calculated through s-parameter measurements, it is essential to understand how it is calculated and evaluate the measurement procedure as TFTs might have specific properties which may make the measurement very different from standard CMOS. The most significant difference for TFTs is that the devices are very resistive. As TFTs are normally biased with a few volts to 20V level and the output current is normally around a few or a few tens µA, the corresponding resistances are around 100kΩ or even a few MΩ level. Resistance at this level is huge particularly in s-parameter measurement as s- parameters are usually normalised with a characteristic resistance of 50Ω or 75Ω. For single port measurement such as S11 and S22, the relation between corresponding impedance(ZL) 3.3 Model Validation 43 (a) Transfer characteristic (b) Output characteristic Fig. 3.9 I-V characteristic of the TFT under test 44 Small Signal Modelling and Sxx(S11/S22) is: Sxx = ZL−Z0 ZL+Z0 (3.19) where Z0 = 50Ω is the characteristic resistance. Here, as ZL is much larger than Z0, the S11orS22 values should be very close to 1 (or very close to 0dB). This fact only changes at higher frequencies when the impedance of the capacitors drops significantly. As discussed be- fore, the frequency range of concern is the range below the cutoff frequency. The impedance of the capacitor at the frequency range should be comparable with the internal components, otherwise, the circuit performance will be dominated by the capacitors. Therefore, the assumption will be valid in the whole frequency range of concern. S11 and S22 are also called reflection coefficient. It expresses the ratio of the reflected power to the incident power of an electromagnetic wave. As the reflection coefficient is close to 1, the transmission coefficients (S12 and S21) should be small as most of the power are reflected rather than transmitted. The experiment shows these two parameters are very close to 0. Now as we know the estimated value of the 4 parameters, we need to evaluate the equation of h21 (Eq. (3.20)). h21 =− 2S21 (1−S11)(1+S22)+S12S21 (3.20) Here, the major complexity comes from the denominator. Therefore, the next question is whether the denominator is dominated by some part of the terms or not. It is pretty hard to tell which part of the denominator is more dominant intuitively, as both terms of (1− S11)(1+ S22) and S12S21 are close to 0. One could estimate from the assumption that 1− S11 has a similar order of magnitude compared with S12 or S21 to conclude that (1−S11)(1+S22) term dominates over S12S21. A further check from experiment based on the IGZO TFTs shows that the former term is 4 orders bigger than the latter term as shown in Fig. 3.10. Therefore, the S12S21 term can be omitted when calculating the current gain. Further, as S22 is very close to 1, the product of 1−S11 and 1+S22 is much more sensitive to the change of S11 than S22. The final approximation of the equation is then as follows: Ai = |h21| ≈ | S21S11−1 | (3.21) Note that the approximation above is very accurate as the dominance of parameters is so significant. A comparison based on experiment result is shown in Fig. 3.11. The estimation gives 0.04% error in calculating the current gain of TFT within the frequency of concern. Therefore, Eq. (3.21) can be used in the measurement of current-gain or fT of TFTs without the measurements of S12 and S22. 3.3 Model Validation 45 Fig. 3.10 Comparison between terms of (1−S11)(1+S22) and S12S21 Now from the derivation and verification above, we know the dominating parameters for TFT fT measurement are S11 and S21. Thus, to evaluate the difference of TFT model and CMOS model, we need to evaluate and analyse the two parameters in both models. As both the two measurements need to connect a matched load at port 2 of the network, the equivalent circuit for the measurements is identical as shown in Fig. 3.12. The modelling of each parameter based on the equivalent circuit will be discussed in the following subsections. And the simulation will be based on the static and dynamic parameters extracted in Chapter 2. S11 measurement and analysis Fig. 3.10 and Fig. 3.11 shows that the measurement at lower frequencies is very noisy especially when S11 is involved. The main reason is the contribution of the 1− S11 term. As discussed before, the S11 parameter is very close to 1 due to the near open-circuit characteristic of capacitors at low frequencies. Therefore, the value of 1−S11 highly depends on the measurement accuracy of S11. According to the datasheet of the network analyser E5061B, the measurement accuracy for S11 with different calibration kit is around 0.03 and 2 degree for magnitude and phase measurement respectively, when average factor equals to 1 and 10Hz bandwidth is selected. However, in order to measure the parameters for the device 46 Small Signal Modelling (a) Calculated h21 (b) Normalized error Fig. 3.11 Comparison between Eq. (3.20) and its approximation Eq. (3.21) 3.3 Model Validation 47 Fig. 3.12 Equivalent circuit for S11 and S21 measurement under test (DUT), the accuracy is not enough as the values of the S11 magnitude and phase are 1 order smaller than the corresponding accuracies. Further, after checking the measurement results, we noticed that the major uncertainty comes from the shift of the machine property over time. As this shift in rather slow compared with one cycle of a frequency sweep, we choose to do two quick sweeps with and without the device connected to measure the S11 value. The open-circuit values are then subtracted from the measurement results to cancel the property shift of the machine. The measurement and simulation results are shown in Fig. 3.13. As can be seen, TFT model fits much better compared with adapted CMOS model. The major difference comes from the magnitude measurement as the measured curve drops, which suggests that there’s at least an extra pole for the S11 expression of TFT model higher but close to the frequency of concern. The calculation based on CMOS model gives the S11 model below: S11CMOS ≈− −g20+(COV S+Cch)g0s+COV D(COV S+Cch)s2 g20+(2g0COV D+g0(COV S+Cch))s+COV D(COV S+Cch)s 2 (3.22) where gx = 1rx is the conductance of the corresponding resistor. The conductance terms are used to make the equation looks simpler. An approximation is also used to simplify the equation based on the assumption that g0 >> gm and g0 >> go. This is particularly true in the case of TFT because of the high resistance of TFT devices. For example, 1/gm is normally around a few kΩ to a few 100kΩ and ro is around a few MΩ. However, the characteristic resistance is 50Ω. Therefore the conductance of the characteristic resistance g0 should be more dominant compared with all other resistances and conductances in the 48 Small Signal Modelling (a) Magnitude (b) Phase Fig. 3.13 Magnitude and phase measurement and simulation for S11 parameter based on TFT model and adapted CMOS model 3.3 Model Validation 49 small signal model of TFT (or the adapted CMOS one). The assumption is also used in the derivation of S11 of TFT model. This model suggests that the equation has 2 zeros and 2 poles. Calculation based on Eq. (3.22) shows that the 2 zeros are at 3.4462GHz and 10.853GHz and 2 poles at 2.4898GHz and 15.022GHz. As the zeros and poles are all far above the frequency of concern (i.e. below 10MHz), the S11 modelled by the CMOS model would remain at 1 at the frequency range as suggested in Fig. 3.13a (which suggests a very close to open circuit behaviour). The calculation based on the equivalent circuit shown in Fig. 3.12 gives the S11 expression as follows: S11T FT =− −gtotgcg20+a1s+a2s2+(gc+goi)CchCOV SCOV Ds3 gtotgcg20+b1s+b2s 2+(gc+goi)CchCOV SCOV Ds3 , (3.23) where gtot = gmi+gc+2goi a1 = [gcgtotCOV S+gc(gc+goi)Cch]g0− (gc+goi)g20Cch, a2 = (gc+goi)CchCOV Sg0, b1 = [gcgtotCOV S+2gtotgcCOV D+(gc+3goi)gcCch]g0+(gc+goi)Cchg20, b2 = (gc+goi)g0Cch(COV S+2COV D). Here, the equation has 3 zeros and 3 poles. To further evaluate the difference between the two models we simulated their S11 parameters. The simulation results are shown in Fig. 3.14. Just as the figure suggests, two zeros lie between two poles to make a valley of the magnitude curves for both models. However, these poles and zeros are at much higher frequencies compared the cutoff of the transistor. Therefore, the simulation results only show a flat line below 10MHz for CMOS model. As for the TFT model, the numeric calculation shows that there are 3 poles at 54.755MHz, 2.8314GHz and 19.389GHz. The zeros are at 54.947MHz, 4.5660GHz and 11.981GHz. The 2 zeros and 2 poles at GHz level have similar values compared with the CMOS model to make a similar shape of the curve at very high frequency. However, they would not create much difference at the low frequency range of our concern either. The major difference we’ve noticed is the extra zero and pole at 54.947MHz and 54.755MHz, respectively. These two points are pretty close to the cutoff frequency and the frequency of concern. From further inspection of Fig. 3.13 and Fig. 3.14, we notice that it is these pole and zero that cause a slight drop of the S11 magnitude. As the pole is slightly smaller than the zero (although only by less than 1%), the magnitude starts to drop a little before the effect is cancelled by the zero. This drop distinguishes the two model at around 10MHz range as the measurement in Fig. 3.13 suggests. This also suggests that the TFT 50 Small Signal Modelling Fig. 3.14 S11 simulation of the TFT and CMOS models model (with the contact resistance and the separation of the channel and overlap capacitance) captures an extra pole and zero, which results in a better accuracy than the CMOS model. It can be noted that |S11| equals to 0dB (or 1) at both very low frequencies and very high frequencies. This can be explained by the connection of the COV S as it acts as open-circuit at low frequencies (S11 = 1) and short-circuit (S11 =−1) at high frequencies. Here, as the major accuracy improvement comes from the extra zero and pole, it is worth finding the exact expression of their analytic form. The accurate expressions can be directly extracted from the Eq. (3.23) by solving the cubic equations. However, the expressions become too long for the purpose of estimating the positions of the extra zero and pole. Alternatively, we discovered that by taking out the Cch in CMOS model, the simulation results of the two models become very similar at high frequencies (shown in Fig. 3.14). Therefore, we could compare the coefficients of Eq. (3.23) and Eq. (3.22) (taken out the Cch term) to find the estimated expression of the extra zero and pole assuming that the CMOS model contains all the poles and zeros except the extra ones. Considering the matching of coefficients, we found the estimated expressions as: 3.3 Model Validation 51  fpe =− gtotgcg02π(g0+gc)(gc+goi)Cch fze =− gtotgcg02π(g0−gc)(gc+goi)Cch (3.24) where fpe is the frequency of the extra pole and fze is the frequency of the extra zero. Here, the expressions give a rather accurate estimation of the extra pole and zero of 54.763MHz and 54.946MHz, respectively, while the numeric calculation gives 54.755MHz and 54.947MHz. Here, with the approximation of the extra zero and pole, one can easily draw the S11 curves without simulating of the more complicated equivalent circuit. It also provides support for the possibility of fitting a S11 measurement with one pole and one zero. S21 measurement and analysis As for S21, the simulation and measurement results are shown in Fig. 3.15. The results show that the difference of S21 between models is less significant compared with S11. This is because of the parameter varies much in a log scale axis (dB) and, therefore, the difference is not visibly prominent in comparison with that of S11. For example, the measurement value at 8MHz is |S21| = 2.85×10−3. The modelled value are |S21T FT | = 2.88×10−3 (yielding 1.1% error) and |S21COMS |= 2.69×10−3 (yielding 5.6% error). The analytic expression of S21 in CMOS model is shown as Eq. (3.25): S21CMOS = 2g0(sCOV D−gm) g20+(2COV D+COV S+Cch)g0s+COV D(COV S+Cch)s 2 (3.25) The equation shows that there are two poles and one zero in the model. Calculation based on the equation shows that the zero is at 6.107MHz and the two poles are at 2.490GHz and 15.02GHz. Again, as the poles are all far above the frequency of concern, they would not affect the measurement results below 10MHz. The results that there’s only one dominant zero within the frequency range. The other point that should be noticed is that at very low frequencies Eq. (3.25) yields to: S21CMOS =− 2gm g0 (3.26) Here, the g0 = 1/50Ω−1 is a known value. Therefore, the S21 measurement can directly reflect the small signal gm for the CMOS model. As discussed in section 3.2.1, the TFT and CMOS models are equivalent at low frequencies as the capacitances can be ignored. Therefore, this gm can also be used to extract the gmi for TFT model at lower frequencies. 52 Small Signal Modelling (a) Magnitude (b) Phase Fig. 3.15 Magnitude and phase measurement and simulation for S21 parameter based on TFT model and adapted CMOS model 3.3 Model Validation 53 As for TFT model, the equivalent circuit for the S21 calculation is the same as the one used for S11 as the equivalent connection for the measurements are the same (Fig. 3.12). Hence, we could derive the equation of S21 as below: S21T FT = 2g0 −g2cgmi+(gcgtotCOV D+goigcCch)s+(gc+goi)COV DCchs2 gtotgcg20+ c1s+ c2s 2+(gc+go)CchCOV SCOV Ds3 , (3.27) where, c1 = (gc+goi)Cchg20, c2 = (gc+goi)Cch(COV S+2COV D)g0. Here, the equation shows two zeros and three poles for TFT model. Based on the parameters of the device under test, we have that the zeros are at 5.501MHz and 60.90MHz and that the poles are at 54.76MHz, 2.831GHz and 19.39GHz. Similar to the case for S11, there’s one extra zero and one extra pole at 60.90MHz and 54.76MHz, respectively. These extra zero and pole would also contribute a slight drop on the magnitude of S21 as they are very close together, just as the case for S11. The difference here is that there’s a dominant zero at 5.501MHz which is different to the 6.107MHz zero for the CMOS model. Hence, the difference of the dominant zero contributes to the major difference between the CMOS and TFT model. And the difference caused by the extra zero and pole becomes less prominent. Here, the dominant zero for S21 can be expressed as: fz0T FT = gc (√ (goiCch+gtotCOV D)2+4gmi(gc+goi)CchCOV D−goiCch+gtotCOV D ) 4πCchCOV D(gc+goi) (3.28) whereas the dominant zero for CMOS model can be express as: fz0CMOS = gm 2πCOV D (3.29) While CMOS model gives a simpler solution for the dominant zero for the S21 parameter, it does not count the contribution of the channel capacitance (Cch). The TFT model, however, accurately models the dominant zero with the contribution of Cch but yields a much more complicated equation. Again, at low frequencies Eq. (3.27) yields to: S21T FT = −2gcgmi gtotg0 (3.30) 54 Small Signal Modelling Fig. 3.16 S11 simulation of the TFT and CMOS models Changing the conductance terms into resistance terms and combining Eq. (3.10), we have: S21T FT = −2gmiroi (gmiroirc+2rc+ roi)g0 =−2gm g0 (3.31) This is exactly Eq. (3.26). This confirms that the CMOS and TFT model can be equivalent at low frequencies. fT measurement and analysis As discussed, we can use Eq. (3.21) to calculate the short-circuit current gain of the TFT with enough accuracy. Therefore, we used the measured S11 and S21 to derive the current gain of the TFT under test and then use a linear fitting to get the value of the fT . The measurement results of fT along with simulations based on the TFT and CMOS model are shown in Fig. 3.16. The measurement results show a cutoff frequency of 3.11MHz while the proposed TFT and CMOS models predict cutoff frequencies of 3.14MHz and 2.72MHz, respectively. This yields an error of 1% and 12.5%, respectively. Here, we proved that the TFT model presented here is more accurate to predict the fT compared with the CMOS counterpart. However, the measurement results of |h21| or |Ai| only 3.3 Model Validation 55 became consistent with the model at frequencies around MHz level. At lower frequencies, the measurement became very noisy as illustrated in Fig. 3.11. However, as shown in the analysis of Eq. (3.16), the |Ai| curve should be a straight line at lower frequencies as it is dominated by a single pole at 0Hz. The main reason for the discrepancy is the measurement accuracy of S11, especially at low frequencies. As the value of |S11| is very close to, if not exactly equals to, 0dB at very low frequencies, the noise of the measurement becomes the major contribution to its value. Therefore, as Eq. (3.21) suggested that the accuracy of |Ai| depends on the subtraction of S11 and 1 (0dB), the noise of the measurement at this frequency range would dominate the results of |Ai| and make them unreliable. This brings a concern for the measurement of other TFTs with lower cutoff frequency, which means the frequency of concern is lower, or with smaller size, which means that the capacitances are smaller making the gate-source port behave even closer to open-circuit. Therefore, we should use a fitting method to fit the S11 parameter before using its value to calculate |Ai| for the cases described above. More detailed discussion regarding S11 fitting will be discussed in the next subsection and also be used to measure the VT shift effect in fT measurement. 3.3.3 fT and VT shift In this subsection, we will discuss the effect of the VT shift. As the accuracy of TFT model is proved in the s-parameter and fT measurements, we would focus on the TFT model. VT shift is another important factor for TFTs, as this is one the major concerns in TFT based circuit designs, especially where TFTs’ analogue properties are of concern, for example in RFID tags. For the case of the small signal model, the TFT is used as a voltage controlled current source for the input of the small signal. Therefore, it is imperative to consider the impact of VT shift on small signal model. In order to examine its influence on the small signal behaviour, a constant bias stress measurement is needed to measure both fT and VT of the transistor. However, simultaneous measurement of the two parameters is not practical as the measurements of them need two different connections, i.e. fT measurement needs network analyser and isolation for DC bias and small signal input and output, however, the VT measurement needs I-V sweep to measure the transfer characteristics (which we use KEITHLEY 4200 to measure). Therefore, we need to find an alternative way to measure both parameters other than keep switching between different measurement platform. Here, we decided to use two different constant bias measurements and connect the two parameters through gm as a bridge. The basic idea is that although fT and VT can not be obtained simultaneously, the gm value can be obtained together in the measurements of both. In specific, gm can be obtained through derivative of transfer characteristics as discussed in 56 Small Signal Modelling section 3.1.1. It can also be obtained through low-frequency S21 measurement as discussed in section 3.3.2. Also, as VT directly affects the value of gm (gmi is considered to be near linearly related to the overdrive voltage (VGS−VT )) and gm also relates with fT , gm can be a perfect link between the two measurements. fT and gm As discussed in section 3.3.2, a fitting method should be applied to the S11 data to obtain a better low-frequency response for the current gain result. Here, as the device under test is smaller in channel width and the cut-off frequency is lower, it is necessary to use the method for better accuracy. The measurement result of S11 is shown in Fig. 3.17. The results are fitted using 1 zero and 1 pole approximation following Eq. (3.32): S11 ≈ 1+as1+bs (3.32) where a and b are fitting parameters. The measurement results are fitted using MATLAB. Here, as illustrated in Fig. 3.17, the measurement results are not reliable below 1MHz for magnitude measurement and below 100kHz for phase measurement. The major cause of the unreliable results is that the value of |S11| or its phase drops below the noise level at low frequency range. Therefore, the fitting method is imperative to obtain a reasonable low-frequency response. In addition, as the cut-off frequency for the device under test in this subsection is around 1.5MHz, which is close to the unreliable range of |S11| (below 1MHz), fitting method is even more important for a better accuracy, especially for capturing the small change of fT caused by VT shift. In order to capture the fT change under stress measurement, we kept the bias stress for 8 hours and measure the corresponding s-parameters. The converted |h21| with and without the fitting of S11 are shown in Fig. 3.18. As illustrated, the |h21| values without fitting are rather random at lower frequencies, which cannot reliably capture the change of the curves versus stress time. However, with fitted S11, we can clearly capture the drop of the cut-off frequency when the stress time increases, as illustrated in the inset of Fig. 3.18b. The low-frequency measurement of S21 can be used to extract gm following a derivation from Eq. (3.31): gm = g0|S21| 2 (3.33) Here, we can use the flat part of the |S21| measurement and calculate the average of |S21| as shown in Fig. 3.19. The value of |S21| drops when bias time increases, which is consistent with the drop of gm caused by the VT shift. 3.3 Model Validation 57 (a) Magnitude (b) Phase Fig. 3.17 Magnitude and phase measurement and fitting for S11 parameter based on single pole and zero approximation 58 Small Signal Modelling (a) with S11 fitting (b) without S11 fitting Fig. 3.18 h21 results with and without S11 fitting 3.3 Model Validation 59 Fig. 3.19 S21 measurement showing low frequency range for gm extraction By extracting gm through Fig. 3.19 and the corresponding fT through Fig. 3.18b, we can get the fT vs. gm relation as shown in Fig. 3.20. The results show a linear approximation to Eq. (3.17). gm and VT The gm vs. VT relation can be measured through I-V measurement of the device. The TFT under test was biased at VGS = 8V and VDS = 15V and I-V sweep was done every logarithmic time interval in order to extract the gm and VT . The constant stress bias is disrupted for a short time due to the measurement. However, because the sweep time (VT measurement) is much shorter compared with the bias time, we assume that the short pause for the constant stress can be omitted. Notwithstanding this, the results showed that VT continued rising after the sweep and the extracted value is then used to extract the relation between gm and ∆VT . The extracted results are illustrated in Fig. 3.21. Combining the results of Fig. 3.20 and Fig. 3.21, we can get the fT vs. ∆VT relation as illustrated in Fig. 3.22. The corresponding VT was obtained through cubic polynomial fitting of the curve in Fig. 3.21. The error bar here is the standard deviation of the fitting to obtain the corresponding result. In specific, the standard deviation of ∆ fT was obtained from the 60 Small Signal Modelling Fig. 3.20 fT vs. gm relation through S-parameter measurement Fig. 3.21 gm vs. VT shift relation through stress measurement with I-V sweep 3.4 Summary and Discussion 61 Fig. 3.22 Extracted ∆ fT vs. VT shift relation. The error bar here is the standard deviation of the fitting to obtain the corresponding result. linear fitting to obtain the corresponding fT . The standard deviation of ∆VT was obtained from the polynomial fitting of the curve in Fig. 3.21. The ∆VT increased to +0.2V in total after the 10 hours bias stress and the respective ∆ fT drops to 0.08MHz almost linearly. This is due to the relatively small ∆VT compared with VGS which leads to a first-order approximation of Eq. (3.17). From Eq. (3.13) & (3.17), we have: ∆ fT ≈− roi(2RC + roi)(COV +Cch)2π (gmiCOV RCroi+(2RC + roi)(COV +Cch))K ′∆VT = β∆VT (3.34) where β is the constant found to be about−3.2×106[Hz/V ] for the examined TFT. Eq. (3.34) allows estimating the shift in unity gain frequency with threshold voltage shift. Here, the coefficient β will be different for different material- and process-based TFTs. 3.4 Summary and Discussion Small signal models are indispensable for the design of analogue circuits, especially in the former for the correct phase margin and bandwidth of amplifiers and analogue filters. This chapter reports on an accurate small signal model that takes into account the contact 62 Small Signal Modelling Table 3.1 Summary of parameters for the CMOS and TFT models Category RC ≈ 0 Cch < 2β1+1. This is very useful when a negative feedback is needed from the power supply. As the circuit requires all TFTs to work in saturation region, we could calculate its working range. The lower limit of the working range is when VDD is just enough to turn on all the TFTs at the TFT ladder. Thus, we have VDD > 3VT . The upper limit of the working range is limited mainly the working condition of T5. As Vout remains a constant, high VDD would eventually increase the gate voltage of T5 making it work in linear region. Therefore, by writing down the saturation condition of T5, we can get the expression of the whole working region: 3 ·VT 2β1+1 (thus a inverse related Vout with respect to VDD), we have: 3 ·VT Vo−VT , and V2 >V1−VT (4.8) By calculating the values of V1 and V2, we have the allowable output voltage range as: Vi+ 1 2 Vre f − 12(VDD+VT )>VDS. Therefore, to maintain a good linearity, the VGS of the TFT should be fixed with a large value (i.e. independent of VDD and with a value higher than the voltage at the source of T1). Among all the nodes in the circuit shown in Fig. 4.6, the gate voltage of T4 meets the aforementioned requirements. Therefore, the circuit is redesigned as shown in Fig. 4.7. The resistor is replaced with TR whose gate electrode is connected to the gate of T4. The size of TR should be chosen to exhibit a resistance of 600kΩ between drain and source to match the output current with the previous design. The calculation result shows that the width of TR should be around 29µm. The simulation of the circuit is done with different WR which is shown in Fig. 4.8. As illustrated, the circuit still works with a low dependence on VDD. However, the performance of this circuit is obviously worse than the one with a real resistor. The main reason of this comes from the small change of T4’s gate voltage (VG4). As the VDD increases, VG4 also 74 TFT analogue circuit building blocks Fig. 4.8 Simulation results of the current reference circuit without resistor where WR is simulated with different value to exhibt the change of output current increases a little due to the none ideality of the devices (i.e. non-infinity output resistance). Therefore, the resistance of TR will decrease according to Eq. 4.16. This will increase the output current and thus forms a positive feedback which deteriorate the circuit performance. In the case where, the performance of the circuit in Fig. 4.7 is accurate enough we should break the positive feedback that appears on TR. The voltage reference circuit developed earlier could be used. Thus, the gate voltage of TR can be more stable, thereby, break the positive feedback loop. The improved circuit schematic is shown in Fig. 4.9. Here, the voltage reference circuit is connected to the gate of TR making the resistance of TR more reliable. In fact, as Eq. 4.2 suggests, we could increase β2 to be higher than 2β1+1 making the output of the voltage inversely related with VDD. Therefore the effect of non-infinity output resistance could be compensated. Hence, a negative feedback can be formed yielding a even better performance. The simulation results of the improved circuit is shown in Fig. 4.10. As can be seen, output current is much more stable than the previous one and even more stable that the one with a resistor. As the negative feedback exists in the new circuit, we have a few more bonuses. One of them is the sensitivity on the value of WR. As can be seen from Fig. 4.10, even sweeping WR at 100µm range, the output currents are all stable and with only 1µA change, which in the previous circuit was achieved with only a few µm change in WR. Therefore, the improved circuit is more stable against process variations on the size of TR. 4.3 Building Blocks and Simulations 75 Fig. 4.9 Improved current reference circuit with negative feedback on TR Fig. 4.10 Simulation results of the improved current reference circuit where WR is simulated with different value to exhibt the change of output current Another bonus is the VT shift resistance. Although most blocks designed in this section are VT shift independent, it is not so for the current reference circuit as T1 and T2 are having different VGS. Therefore, as operation time increases, T1 will become more conductive compared with T2 and thus increasing the output current (by effectively reducing the value of 76 TFT analogue circuit building blocks Fig. 4.11 Differential amplifier stage with pseudo p-type mirror K2 K1 in Eq. 4.13). However, the VT shift of TR will also increase its resistance thus reducing the output current. Negative feedback also exists in terms of VT shift. Therefore, this more complicated version of current reference yields better VDD independence and at the same time better VT shift resistance. 4.3.4 Differential Amplifier In this section, we consider designing an amplifying stage. With the current mirror designed in section 4.3.2, we could replicate a CMOS differential stage by replacing the p-type current mirror. The design is shown in Fig. 4.11. Here, Vre f is connected to ground to optimize the output range. As the pseudo p-type mirror has big output resistance, the voltage gain of this differential stage should be the same as that of a real p-type mirror, in the mean time, current steering feature would also be available thus benefit the gain by an additional factor of 2. The sizes used for simulation is W3,4 = 50µm, W1,2 = 500µm and W5 = 100µm. T1 and T2 are designed with big W/L ratio for a higher gm. Gate voltage of T5 is set with 3V bias for a 1.5µA bias current on T5 which matches the design of the current reference circuit in the previous section. The transient and ac simulation results are shown in Fig. 4.12. The results suggest that proposed gain stage have over 40dB or 100 gain within the 3dB bandwidth. This is close to 4.3 Building Blocks and Simulations 77 (a) Transient simulation results of the proposed differential stage with 1mV pp input at 1kHz for 5ms (b) Gain-phase simulation of the proposed differential stage yielding a 600kHz unit-gain bandwidth Fig. 4.12 Simulation results of the proposed differential amplifier stage 78 TFT analogue circuit building blocks Fig. 4.13 Full op-amp design with supply independent bias the intrinsic gain of the TFT at the working condition demonstrating that the pseudo p-type current mirror can be used as an promising load in fully n-type circuit design. The gain-phase simulation suggests a 600kHz unit-gain bandwidth of the stage. 4.4 Full Op-Amp Design Combining all the building blocks we designed earlier and adding an output stage (which is a source follower stage), we have the full design of a op-amp which is illustrated in Fig. 4.13. The output stage is added to improve the driving capability of the circuit by reducing the output resistance. Simulation result of the proposed op-amp with open circuit load is illustrated in Fig. 4.14. The op-amp here exhibit a gain of 40dB which benefits from the high-gain of the differential stage and a unit-gain bandwidth of 300kHz, slightly reduced compared with the differential pair due to the added stage. The performance here is enough for many bio-medical applications where signal frequencies are below a few 10kHz. The output resistance is reduced to 50kΩ making it capable of driving many CMOS circuits. 4.5 Current Mirror Layout and Measurement In order to evaluate the effectiveness of the circuit design technique, Indium-Silicon-Oxide (ISO) based current mirrors are fabricated under vacuum process with different layouts. As will be noted, the layout design could also significantly affect the circuit performance. 4.5 Current Mirror Layout and Measurement 79 (a) Transient simulation results of the proposed op-amp with 1mV pp input at 1kHz for 5ms (b) Gain-phase simulation of the proposed op-amp yielding a 300kHz unit-gain bandwidth Fig. 4.14 Simulation results of the proposed op-amp 80 TFT analogue circuit building blocks Fig. 4.15 Schematic of the fabricated ISO current mirror. The circuit schematic is illustrated in Fig. 4.15. Assuming a perfect match between the two transistors, the drain current of T2 (IDS2) would copy the current that flows through the drain of T1 (IDS1). However, mismatch can happen due to the variance of material properties and dimensions in the fabrication process. Therefore, we designed two different layout for the current mirror. The first one is based on two separate devices which is shown in Fig. 4.16a. As the two devices are located separately, device mismatch would totally be reflected on the output current. The current mirror with separate device layout is measured sweeping the voltage supply at the drain of T1 (VDS1 = VGS) where the drain voltage of T2 (VDS2) is kept at 10V. The measurement result is illustrated in Fig. 4.16b. As can be seen, the current of the two transistors are rising according to the increase of VGS. However, there is a mismatch in current as IDS2 increases slower than IDS1. The normalized current error is between 16% and 32% at a VGS range of 5V to 10V, which is illustrated in the inset of Fig. 4.16b. The other layout is designed based on the interdigitated stacked layout. The fabricated layout is shown in Fig. 4.17a. The layout effectively separate each device into 5 devices connected in parallel. This would average out mismatches between devices and improve the matching of the overall device. Also, devices are placed adjacent to each other in a interdigitated way to reduce the mismatch over long distance. The same measurement is done for a current mirror using the interdigitated stacked layout. The measurement result is illustrated in Fig. 4.17b where VDS2 is maintained at 10V. As can be seen, the matching of the transistors is significantly improved. The normalized current error has been reduced to below 8% at the VGS range of 5 to 10V. This indicates that the 4.5 Current Mirror Layout and Measurement 81 (a) Fabricated current mirror under microscope (b) Measurement of the fabricated current mirror Fig. 4.16 Picture and measurement result of the fabricated current mirror based on a separate- device layout. 82 TFT analogue circuit building blocks (a) Fabricated current mirror under microscope (b) Measurement of the fabricated current mirror Fig. 4.17 Picture and measurement result of the fabricated current mirror based on a interdig- itated stacked layout. 4.6 Summary and Discussion 83 layout technique used in silicon CMOS technology can be adapted to the vacuum processed TFTs. Due to the large mismatch ( 20%) of separate devices, to fabricate a circuit based on dimension ratios reported in this chapter is very difficult without a specific layout design. It maybe beneficial to fabricate every device in a multi-finger layout. However, here we consider the full layout design as a future extension of this dissertation. 4.6 Summary and Discussion Op-amp is a fundamental building block for any analogue systems. This chapter delves into the fully n-type circuits with a goal of building a bridge between the underdeveloped fully n-type TFT circuits and the well developed CMOS circuits. The result here is fruitful. With the pseudo p-type current mirror, many CMOS design can be converted to fully n-type designs with comparable performances. The VT independent feature of most building blocks here also simplifies the VT shift compensation procedure where a clock based compensation system is not needed. One noticeable drawback of the pseudo p-type mirror is that the output range is rather limited compared with the PMOS counterpart. An external reference voltage supply may be needed in the case when a very wide output range is required. A dynamic reference could also be designed if another circuit block is added to generate a dynamic Vre f . In terms of power consumption, the feedback unit adds two more current path which consumes extra power. Designers should reduce the size of the TFTs used in this block to minimize the current and hence minimize the power consumption. The op-amp designed here is a simple version to demonstrate the functionality of all the building blocks. More stages maybe needed depending on the requirement of applications. The high gain of the differential stage would effectively reduce the input referred noise making the existing blocks promising in all kinds of potential applications. The most outstanding issue remains here is the rail-to-rail output stage. As in this aspect integrated op-amps mostly benefit from the push-pull architecture which is enabled by complementary devices, design an output stage with similar performance remains a challenging topic with fully n-type devices. Although the design reported in this chapter are all based on n-type devices, they are also usable in fully p-type designs for example in OTFTs due to the symmetrical nature of the two types of devices. The fabrication of the circuit would need a lot of tuning of the layout and models regarding the variation of parameters. Here, due to the mismatch and especially the difficulty 84 TFT analogue circuit building blocks of controlling the value of VT , the full circuit fabrication is considered as a future extension of the research reported in this thesis. We have shown a current mirror stage to illustrate that 20% error is likely to happen between any two separate TFTs on the same substrate. Although the matching can be improved to some extent using layout techniques, special layout is needed for the dimension ratio based circuit designs. Therefore, extensive work is necessary to achieve a reliable circuit performance. Chapter 5 Conclusion and Future Work With the ultimate goal of creating circuits and systems with more functionality, this dis- sertation has presented the whole process of designing circuits with IGZO TFTs starting with accurate DC and AC modelling all the way to circuit designs taking into account the interaction between specific device performances and corresponding alternation of designs. Physical and empirical models of IGZO TFTs have been successfully implemented in the Cadence simulation environment with good smoothness and convergence for fast circuit simulation. A more precise small signal model for TFTs has been developed taken into account the non-idealities of TFTs. The proposed model is tested on IGZO TFTs and is potentially able to work for TFTs from other material families. Because the effects being considered generally exist in most material families due to the amorphous nature of semiconductors used in TFTs, although the underlying physics can be different. The proposed model yields much better accuracy in comparison with widely adapted CMOS small signal model. Further, we discussed in a general way how TFT performance could affect circuit design from a device-circuit interaction stand point. Geometric and temperature dependence of TFT current accuracy are characterised through measurement, statistics and modelling. Techniques for extraction of defects and ageing in devices using closed-loop feedback are discussed with their possible extension to other applications. Finally, fundamental building blocks including voltage and current reference circuits, pseudo p-type current mirror and differential gain stages are designed using n-type IGZO TFTs and simulated through the model developed in this dissertation. The op-amp designed with the aforementioned building blocks are presented for further application-orientated engineering. As the dissertation mainly contributes to the device modelling, simulation and mono-type TFT circuit design, several extensions could be done in the future. 86 Conclusion and Future Work 1. On device modelling, one noticeable improvement could be adding the VT shift model. This is particularly difficult because the simulation environment would expect a device performance to be stable especially in AC simulations. More importantly, further study should be done physically and mathematically on the time-dependent (or real time) VT model. Although we have mentioned in several chapters the stretched exponential function of the VT shift, the function has fundamental flaws in circuit simulations due to the untraceable initial conditions. For example, if we switch on and off a transistor several times, all the parameters in the stretched exponential function (i.e. β and τ) would and should change. However, modelling β and τ based on switching events is very difficult if not impossible. Preliminary work along the line have been reported [? ], although there is room for refinement if thermodynamic considerations are brought into the picture. 2. Several other circuit building blocks could be designed to expand the library of ana- logue building blocks. One of which is a rail-to-rail output stage. As class AB amplifier needs complementary devices to work, it will be meaningful to achieve a similar performance with mono-type TFTs. 3. The temperature dependence of circuit blocks was not compensated in the designs presented. Therefore, another meaningful improvement is to design a temperature independent circuit, in particular, a temperature independent reference circuit. References [1] A. Nathan, A. Ahnood, M. T. Cole, S. Lee, Y. Suzuki, P. Hiralal, F. Bonaccorso, T. Hasan, L. Garcia-Gancedo, A. Dyadyusha, S. Haque, P. Andrew, S. Hofmann, J. Moultrie, D. Chu, A. J. Flewitt, A. C. Ferrari, M. J. Kelly, J. Robertson, G. A. J. Amaratunga, and W. I. Milne, “Flexible electronics: The next ubiquitous platform,” in Proceedings of the IEEE, vol. 100, pp. 1486–1517, 2012. [2] F.-g. Structure, T. Yokota, T. Sekitani, T. Tokuhara, N. Take, U. Zschieschang, M. Takamiya, T. Sakurai, and T. Someya, “Sheet-Type Flexible Organic Active Matrix Amplifier System Using Pseudo-CMOS Circuits With,” vol. 59, no. 12, pp. 3434–3441, 2012. [3] H. Marien, M. S. J. Steyaert, E. Van Veenendaal, and P. Heremans, “A fully integrated sigma-delta ADC in organic thin-film transistor technology on flexible plastic foil,” in IEEE Journal of Solid-State Circuits, vol. 46, pp. 276–284, 2011. [4] L. Zhou, A. Wanga, S.-C. Wu, J. Sun, S. Park, and T. N. Jackson, “All-organic active matrix flexible display,” Applied Physics Letters, vol. 88, p. 083502, feb 2006. [5] D.-U. Jin, J.-S. Lee, T.-W. Kim, S.-G. An, D. Straykhilev, Y.-S. Pyo, H.-S. Kim, D.-B. Lee, Y.-G. Mo, H.-D. Kim, and H.-K. Chung, “65.2: Distinguished Paper: World-Largest (6.5”) Flexible Full Color Top Emission AMOLED Display on Plastic Film and Its Bending Properties,” SID Symposium Digest of Technical Papers, vol. 40, no. 1, p. 983, 2009. [6] S.-H. K. Park, M. Ryu, C.-S. Hwang, S. Yang, C. Byun, J.-I. Lee, J. Shin, S. M. Yoon, H. Y. Chu, K. I. Cho, K. Lee, M. S. Oh, and S. Im, “42.3: Transparent ZnO Thin Film Transistor for the Application of High Aperture Ratio Bottom Emission AM-OLED Display,” SID Symposium Digest of Technical Papers, vol. 39, no. 1, p. 629, 2008. [7] C. Zysset, N. Munzenrieder, L. Petti, L. Buthe, G. A. Salvatore, and G. Troster, “IGZO TFT-Based All-Enhancement Operational Amplifier Bent to a Radius of 5 mm,” IEEE Electron Device Letters, vol. 34, pp. 1394–1396, nov 2013. [8] T. Tanaka, H. Asuma, K. Ogawa, Y. Shinagawa, K. Ono, and N. Konishi, “An LCD addressed by a-Si:H TFTs with peripheral poly-Si TFT circuits,” Proceedings of IEEE International Electron Devices Meeting, 1993. [9] K. Nomura, H. Ohta, A. Takagi, T. Kamiya, M. Hirano, and H. Hosono, “Room- temperature fabrication of transparent flexible thin-film transistors using amorphous oxide semiconductors.,” Nature, vol. 432, pp. 488–492, 2004. 88 References [10] D. Gundlach, Y. Lin, T. Jackson, S. Nelson, and D. Schlom, “Pentacene organic thin-film transistors-molecular ordering and mobility,” IEEE Electron Device Letters, vol. 18, 1997. [11] R. M. A. Dawson and M. G. Kane, “24.1: Invited Paper: Pursuit of Active Matrix Organic Light Emitting Diode Displays,” SID Symposium Digest of Technical Papers, vol. 32, no. 1, p. 372, 2001. [12] V. Fiore, E. Ragonese, S. Abdinia, S. Jacob, I. Chartier, R. Coppard, A. Van Roermund, E. Cantatore, and G. Palmisano, “30.4 A 13.56MHz RFID tag with active envelope detection in an organic complementary TFT technology,” in Digest of Technical Papers - IEEE International Solid-State Circuits Conference, vol. 57, pp. 492–493, 2014. [13] G. Maiellaro, E. Ragonese, R. Gwoziecki, S. Jacobs, N. Marjanovic, M. Chrapa, J. Schleuniger, and G. Palmisano, “Ambient light organic sensor in a printed com- plementary organic TFT technology on flexible plastic foil,” IEEE Transactions on Circuits and Systems I: Regular Papers, vol. 61, pp. 1036–1043, 2014. [14] K. Y. Chung and G. W. Neudeck, “Transient analysis of the CMOS-like a-SI:H TFT inverter circuit,” IEEE Journal of Solid-State Circuits, vol. 24, pp. 822–829, 1989. [15] C. Zysset, N. Munzenrieder, T. Kinkeldei, K. Cherenack, and G. Troster, “Indium- gallium-zinc-oxide based mechanically flexible transimpedance amplifier,” Electronics Letters, vol. 47, no. 12, p. 691, 2011. [16] T. Brody, “The thin film transistor-A late flowering bloom,” IEEE Transactions on Electron Devices, vol. 31, pp. 1614–1628, nov 1984. [17] S. Lee and A. Nathan, “Subthreshold Schottky-barrier thin-film transistors with ul- tralow power and high intrinsic gain,” Science, vol. 354, no. 6310, pp. 302–304, 2016. [18] M. S. Shur, “SPICE Models for Amorphous Silicon and Polysilicon Thin Film Tran- sistors,” Journal of The Electrochemical Society, vol. 144, no. 8, p. 2833, 1997. [19] R. Chaji and A. Nathan, Thin Film Transistor Circuits and Systems. Cambridge University Press, 2013. [20] S. Lee, S. Jeon, R. Chaji, and A. Nathan, “Transparent semiconducting oxide technol- ogy for touch free interactive flexible displays,” Proceedings of the IEEE, vol. 103, no. 4, pp. 644–664, 2015. [21] S. Jeon, I. Song, S. Lee, B. Ryu, S. E. Ahn, E. Lee, Y. Kim, A. Nathan, J. Robertson, and U. I. Chung, “Origin of high photoconductive gain in fully transparent hetero- junction nanocrystalline oxide image sensors and interconnects,” Advanced Materials, vol. 26, no. 41, pp. 7102–7109, 2014. [22] A. Nathan, S. Lee, S. Jeon, I. Song, and U.-i. Chung, “Amorphous Oxide TFTs: Progress and Issues,” SID Symposium Digest of Technical Papers information display, vol. 43, no. 1, pp. 1–4, 2012. References 89 [23] H. Aoki, “Dynamic characterization of a-Si TFT-LCD pixels,” IEEE Transactions on Electron Devices, vol. 43, no. 1, pp. 31–39, 1996. [24] K. Karim, P. Servati, N. Mohan, A. Nathan, and J. Rowlands, “VHDL-AMS modeling and simulation of a passive pixel sensor in a-Si:H technology for medical imaging,” in ISCAS 2001. The 2001 IEEE International Symposium on Circuits and Systems (Cat. No.01CH37196), vol. 5, pp. 479–482, IEEE. [25] K. Khakzar and E. Lueder, “Modeling of amorphous-silicon thin-film transistors for circuit simulations with SPICE,” IEEE Transactions on Electron Devices, vol. 39, pp. 1428–1434, jun 1992. [26] L. Resendiz, B. Iniguez, a. Estrada, and a. Cerdeira, “Modification of amorphous level 15 AIM SPICE model to include new subthreshold model,” 2004 24th International Conference on Microelectronics (IEEE Cat. No.04TH8716), vol. 1, pp. 16–19, 2004. [27] T. Leroux, “Static and dynamic analysis of amorphous-silicon field-effect transistors,” Solid-State Electronics, vol. 29, pp. 47–58, jan 1986. [28] M. Shur, M. Hack, and J. G. Shaw, “A new analytic model for amorphous silicon thin-film transistors,” Journal of Applied Physics, vol. 66, no. 7, p. 3371, 1989. [29] G. Horowitz and P. Delannoy, “An analytical model for organic-based thin-film transistors,” Journal of Applied Physics, vol. 70, no. 1, p. 469, 1991. [30] M. C. J. M. Vissenberg and M. Matters, “Theory of the field-effect mobility in amorphous organic transistors,” vol. 57, no. 20, p. 13, 1998. [31] M. Estrada, A. Cerdeira, J. Puigdollers, L. Reséndiz, J. Pallares, L. Marsal, C. Voz, and B. Iñiguez, “Accurate modeling and parameter extraction method for organic TFTs,” Solid-State Electronics, vol. 49, pp. 1009–1016, jun 2005. [32] M. Fadlallah, W. Benzarti, G. Billiot, W. Eccleston, and D. Barclay, “Modeling and characterization of organic thin film transistors for circuit design,” Journal of Applied Physics, vol. 99, no. 10, p. 104504, 2006. [33] A. Valletta, A. S. Demirkol, G. Maira, M. Frasca, V. Vinciguerra, L. G. Occhipinti, L. Fortuna, L. Mariucci, and G. Fortunato, “A compact Spice model for organic TFTs and applications to logic circuit design,” in 2015 IEEE 15th International Conference on Nanotechnology (IEEE-NANO), pp. 1434–1437, IEEE, jul 2015. [34] H.-H. Hsieh, T. Kamiya, K. Nomura, H. Hosono, and C.-C. Wu, “Modeling of amorphous InGaZnO[sub 4] thin film transistors and their subgap density of states,” Applied Physics Letters, vol. 92, no. 13, p. 133503, 2008. [35] S. Lee, K. Ghaffarzadeh, A. Nathan, J. Robertson, S. Jeon, C. Kim, I. H. Song, and U. I. Chung, “Trap-limited and percolation conduction mechanisms in amorphous oxide semiconductor thin film transistors,” Applied Physics Letters, vol. 98, 2011. [36] S. Lee, S. Jeon, and A. Nathan, “Modeling Sub-Threshold Current–Voltage Character- istics in Thin Film Transistors,” Journal of Display Technology, vol. 9, pp. 883–889, nov 2013. 90 References [37] S. Lee and A. Nathan, “Localized tail state distribution in amorphous oxide transistors deduced from low temperature measurements,” Applied Physics Letters, vol. 101, no. 11, p. 113502, 2012. [38] S. Lee, D. Striakhilev, S. Jeon, and A. Nathan, “Unified Analytic Model for Cur- rent–Voltage Behavior in Amorphous Oxide Semiconductor TFTs,” IEEE Electron Device Letters, vol. 35, pp. 84–86, jan 2014. [39] Z. Zong, L. Li, J. Jang, N. Lu, and M. Liu, “Analytical surface-potential compact model for amorphous-IGZO thin-film transistors,” Journal of Applied Physics, vol. 117, no. 21, p. 215705, 2015. [40] C. Perumal, K. Ishida, R. Shabanpour, B. K. Boroujeni, L. Petti, N. S. Munzenrieder, G. A. Salvatore, C. Carta, G. Troster, and F. Ellinger, “A compact a-IGZO TFT model based on MOSFET SPICE Level=3 template for Analog/RF circuit designs,” IEEE Electron Device Letters, vol. 34, no. 11, pp. 1391–1393, 2013. [41] A. a. Fomani and A. Nathan, “Metastability mechanisms in thin film transistors quantitatively resolved using post-stress relaxation of threshold voltage,” Journal of Applied Physics, vol. 109, no. 8, p. 084521, 2011. [42] G. Horowitz, “Organic Field-Effect Transistors,” Advanced Materials, vol. 10, pp. 365– 377, mar 1998. [43] C. H. Kim, A. Castro-Carranza, M. Estrada, A. Estrada, Y. Horowitz, G. Horowitz, and B. Iniguez, “A compact model for organic field-effect transistors with improved output asymptotic behaviors,” IEEE Transactions on Electron Devices, vol. 60, no. 3, pp. 1136–1141, 2013. [44] L. L. L. Li, H. Marien, J. Genoe, M. Steyaert, and P. Heremans, “Compact Model for Organic Thin-Film Transistor,” IEEE Electron Device Letters, vol. 31, no. 3, pp. 210–212, 2010. [45] S. Lee, Y. Yang, and A. Nathan, “Modeling Above-Threshold Current-Voltage Char- acteristics in Oxide TFTs for Display Pixel Circuits,” Digest of Technical papers, International Symposium, Display Week 2013, SID (Society for Information Display) Symposium, vol. 44, no. 1, pp. 22–25, 2013. [46] “IEEE Recommended Practices #P1485 on: Test Procedures for Micro-electronic MOSFET Circuit Simulator Model Validation.” http://electronica.ugr.es/˜amroldan/modulos/investiga/material_restringido/estandar_medida_mos.pdf, 1997. [47] G. H. See, X. Zhou, K. Chandrasekaran, S. B. Chiah, Z. Zhu, C. Wei, S. Lin, G. Zhu, and G. H. Lim, “A compact model satisfying Gummel symmetry in higher order derivatives and applicable to asymmetric MOSFETs,” IEEE Transactions on Electron Devices, vol. 55, no. 2, pp. 624–631, 2008. [48] L. W. Nagel and D. O. Pederson, “SPICE (Simulation Program with Integrated Circuit Emphasis),” No ERLM382Electronic Research Laboratory, vol. Memorandum, pp. 1– 4, 1973. References 91 [49] B. Sheu, D. Scharfetter, P.-K. Ko, and M.-C. Jeng, “BSIM: Berkeley short-channel IGFET model for MOS transistors,” 1987. [50] IEEE, Verilog IEEE Std 1364-1995. 1996. [51] H. Verilog, “Verilog-AMS Language Reference Manual,” Eda-Stds.Org, 2014. [52] L. Lemaitre, C. McAndrew, and S. Hamm, “ADMS-automatic device model synthe- sizer,” Proceedings of the IEEE 2002 Custom Integrated Circuits Conference (Cat. No.02CH37285), pp. 27–30, 2002. [53] L. Lemaitre, G. Coram, C. McAndrew, and K. Kundert, “Extensions to Verilog-A to support compact device modeling,” Proceedings of the 2003 IEEE International Workshop on Behavioral Modeling and Simulation, 2003. [54] M. A. Chalkiadaki, C. Valla, F. Poullet, and M. Bucher, “Why- and how- to integrate Verilog-A compact models in SPICE simulators,” International Journal of Circuit Theory and Applications, vol. 41, no. July 2012, pp. 1203–1211, 2013. [55] G. Depeyrot, “Verilog-A Compact Model Coding Whitepaper,” Memory, vol. 2, no. June, pp. 821– 824, 2010. [56] A. S. Sedra and K. C. Smith, Microelectronic circuits. New York Oxford University Press, 6th ed., 2011. [57] M. Powell, “The physics of amorphous-silicon thin-film transistors,” IEEE Transac- tions on Electron Devices, vol. 36, no. 12, pp. 2753–2763, 1989. [58] P. Servati, D. Striakhilev, and A. Nathan, “Above-threshold parameter extraction and modeling for amorphous silicon thin-film transistors,” IEEE Transactions on Electron Devices, vol. 50, pp. 2227–2235, nov 2003. [59] P. V. Necliudov, M. S. Shur, D. J. Gundlach, and T. N. Jackson, “Modeling of organic thin film transistors of different designs,” Journal of Applied Physics, vol. 88, no. 11, p. 6594, 2000. [60] H. Klauk, G. Schmid, W. Radlik, W. Weber, L. S. Zhou, C. D. Sheraw, J. A. Nichols, and T. N. Jackson, “Contact resistance in organic thin film transistors,” Solid-State Electron., vol. 47, pp. 297–301, 2003. [61] P. V. Necliudov, M. S. Shur, D. J. Gundlach, and T. N. Jackson, “Contact resistance extraction in pentacene thin film transistors,” Solid-State Electronics, vol. 47, pp. 259– 262, feb 2003. [62] M. J. Powell, C. van Berkel, and J. R. Hughes, “Time and temperature dependence of instability mechanisms in amorphous silicon thin-film transistors,” Applied Physics Letters, vol. 54, no. 14, p. 1323, 1989. [63] S. J. Zilker, C. Detcheverry, E. Cantatore, and D. M. De Leeuw, “Bias stress in organic thin-film transistors and logic gates,” Applied Physics Letters, vol. 79, no. 8, pp. 1124–1126, 2001. 92 References [64] H.-W. Zan and S.-C. Kao, “The Effects of Drain-Bias on the Threshold Voltage Instability in Organic TFTs,” IEEE Electron Device Letters, vol. 29, pp. 155–157, feb 2008. [65] S. Lee, A. Nathan, S. Jeon, and J. Robertson, “Oxygen Defect-Induced Metastability in Oxide Semiconductors Probed by Gate Pulse Spectroscopy,” Scientific Reports, vol. 5, no. October, p. 14902, 2015. [66] R. B. M. Cross and M. M. De Souza, “Investigating the stability of zinc oxide thin film transistors,” Applied Physics Letters, vol. 89, no. 26, p. 263513, 2006. [67] J. Brittain, “Thevenin’s theorem,” IEEE Spectrum, vol. 27, p. 42, mar 1990. [68] D. H. Johnson, “Origins of the equivalent circuit concept: The current-source equiva- lent,” 2003. [69] S. Sambandan and A. Nathan, “A stable n-channel mirrorable current source for versatile analog design with thin film transistors,” in 48th Midwest Symposium on Circuits and Systems, 2005., pp. 836–839 Vol. 1, IEEE, 2005. [70] L. Toledo, C. Dualibe, P. Petrashin, and W. Lancioni, “A novel supply-independent biasing scheme for use in CMOS voltage reference,” in Proc. Design of Circuits and Integrated Systems (DCIS), (Lisbon, Portugal), 2005. [71] E. Vittoz and J. Fellrath, “CMOS analog integrated circuits based on weak inversion operations,” IEEE Journal of Solid-State Circuits, vol. 12, pp. 224–231, jun 1977. [72] S. Sambandan and A. Nathan, “Equivalent Circuit Description of Threshold Voltage Shift in a-Si:H TFTs From a Probabilistic Analysis of Carrier Population Dynamics,” IEEE Transactions on Electron Devices, vol. 53, pp. 2306–2311, sep 2006. [73] H. Klauk, “Organic thin-film transistors,” Chemical Society Reviews, vol. 39, no. 7, p. 2643, 2010. [74] L. Feng, W. Tang, X. Xu, Q. Cui, and X. Guo, “Ultralow-Voltage Solution-Processed Organic Transistors With Small Gate Dielectric Capacitance,” IEEE Electron Device Letters, vol. 34, pp. 129–131, jan 2013. [75] Jae Sang Lee, Seongpil Chang, Sang-Mo Koo, and Sang Yeol Lee, “High-Performance a-IGZO TFT With ZrO2 Gate Dielectric Fabricated at Room Temperature,” IEEE Electron Device Letters, vol. 31, pp. 225–227, mar 2010. [76] J.-S. Park, J. K. Jeong, Y.-G. Mo, H. D. Kim, and S.-I. Kim, “Improvements in the device characteristics of amorphous indium gallium zinc oxide thin-film transistors by Ar plasma treatment,” Applied Physics Letters, vol. 90, no. 26, p. 262106, 2007. [77] D. Braga and G. Horowitz, “High-Performance organic field-effect transistors,” 2009. [78] X. Guo and S. R. P. Silva, “ENGINEERING: High-Performance Transistors by Design,” Science, vol. 320, pp. 618–619, may 2008. [79] C. Dimitrakopoulos and P. Malenfant, “Organic thin film transistors for large area electronics,” Advanced Materials, vol. 14, no. 2, pp. 99–117, 2002. References 93 [80] S.-J. Kim, S.-Y. Lee, Y. W. Lee, S.-H. Kuk, J.-Y. Kwon, and M.-K. Han, “Effect of Charge Trapping/Detrapping on Threshold Voltage Shift of IGZO TFTs under AC Bias Stress,” Electrochemical and Solid-State Letters, vol. 15, no. 4, p. H108, 2012. [81] A. Suresh and J. F. Muth, “Bias stress stability of indium gallium zinc oxide chan- nel based transparent thin film transistors,” Applied Physics Letters, vol. 92, no. 3, p. 033502, 2008. [82] Y. Qiu, Y. Hu, G. Dong, L. Wang, J. Xie, and Y. Ma, “H[sub 2]O effect on the stability of organic thin-film field-effect transistors,” Applied Physics Letters, vol. 83, no. 8, p. 1644, 2003. [83] K. H. Ji, J.-I. Kim, Y.-G. Mo, J. H. Jeong, S. Yang, C.-S. Hwang, S.-H. K. Park, M.-K. Ryu, S.-Y. Lee, and J. K. Jeong, “Comparative Study on Light-Induced Bias Stress Instability of IGZO Transistors With SiNx and SiO2,” IEEE Electron Device Letters, vol. 31, pp. 1404–1406, dec 2010. [84] H. Marien, M. Steyaert, E. van Veenendaal, and P. Heremans, “Analog techniques for reliable organic circuit design on foil applied to an 18dB single-stage differential amplifier,” Organic Electronics, vol. 11, pp. 1357–1362, aug 2010. [85] S. Jeon, S.-E. Ahn, I. Song, C. J. Kim, U.-I. Chung, E. Lee, I. Yoo, A. Nathan, S. Lee, J. Robertson, and K. Kim, “Gated three-terminal device architecture to eliminate persistent photoconductivity in oxide semiconductor photosensor arrays,” Nature Materials, vol. 11, no. 4, pp. 301–305, 2012. [86] S. Sambandan, “High-gain amplifiers with amorphous-silicon thin-film transistors,” IEEE Electron Device Letters, vol. 29, pp. 882–884, 2008. [87] P. Bahubalindruni, V. G. Tavares, P. G. De Oliveira, P. Barquinha, R. Martins, and E. Fortunato, “High-gain amplifier with n-type transistors,” 2013 IEEE International Conference of Electron Devices and Solid-State Circuits, EDSSC 2013, vol. 2, no. 2, 2013. [88] Y.-C. Tarn, P.-C. Ku, H.-H. Hsieh, and L.-H. Lu, “An Amorphous-Silicon Operational Amplifier and Its Application to a 4-Bit Digital-to-Analog Converter,” IEEE Journal of Solid-State Circuits, vol. 45, pp. 1028–1035, may 2010. [89] X. Cheng, S. Lee, G. Yao, and A. Nathan, “TFT Compact Modeling,” Journal of Display Technology, vol. 12, pp. 898–906, sep 2016. [90] O. Marinov, M. J. Deen, U. Zschieschang, and H. Klauk, “Organic thin-film transistors: Part I-compact DC modeling,” IEEE Transactions on Electron Devices, vol. 56, no. 12, pp. 2952–2961, 2009. [91] G. R. Chaji, S. Alexander, J. Marcel Dionne, Y. Azizi, C. Church, J. Hamer, J. Spindler, and A. Nathan, “Stable RGBW AMOLED display with OLED degradation compen- sation using electrical feedback,” Digest of Technical Papers - IEEE International Solid-State Circuits Conference, vol. 53, no. 3, pp. 118–119, 2010. 94 References [92] J. K. Jeong, J. H. Jeong, J. H. Choi, J. S. Im, S. H. Kim, H. W. Yang, K. N. Kang, K. S. Kim, T. K. Ahn, H.-J. Chung, M. Kim, B. S. Gu, J.-S. Park, Y.-G. Mo, H. D. Kim, and H. K. Chung, “3.1: Distinguished Paper: 12.1-Inch WXGA AMOLED Display Driven by Indium-Gallium-Zinc Oxide TFTs Array,” SID Symposium Digest of Technical Papers, vol. 39, no. 1, p. 1, 2008. [93] J.-H. Lee, W.-J. Nam, S.-H. Jung, and M.-K. Han, “A New Current Scaling Pixel Circuit for AMOLED,” IEEE Electron Device Letters, vol. 25, pp. 280–282, may 2004. [94] M. Bagheri, X. Cheng, J. Zhang, S. Lee, S. Ashtiani, and A. Nathan, “Threshold Voltage Compensation Error in Voltage Programmed AMOLED Displays,” Journal of Display Technology, vol. 12, pp. 658–664, jun 2016. [95] S. Lee and A. Nathan, “Conduction Threshold in Accumulation-Mode InGaZnO Thin Film Transistors,” Scientific Reports, vol. 6, no. October 2015, p. 22567, 2016. [96] A. Nathan, G. R. Chaji, and S. J. Ashtiani, “Driving schemes for a-Si and LTPS AMOLED displays,” IEEE/OSA Journal of Display Technology, vol. 1, no. 2, pp. 267– 277, 2005. [97] Y. Lin, H.-P. Shieh, and J. Kanicki, “A Novel Current-Scaling a-Si:H TFTs Pixel Electrode Circuit for AM-OLEDs,” IEEE Transactions on Electron Devices, vol. 52, pp. 1123–1131, jun 2005. [98] Y.-C. Lin and H.-P. Shieh, “A Novel Current Memory Circuit for AMOLEDs,” IEEE Transactions on Electron Devices, vol. 51, pp. 1037–1040, jun 2004. [99] M. Ohta, H. Tsutsu, H. Takahara, I. Kobayashi, T. Uemura, and Y. Takubo, “9.4: A Novel Current Programmed Pixel for Active Matrix OLED Displays,” SID Symposium Digest of Technical Papers, vol. 34, no. 1, p. 108, 2003. [100] T. Sasaoka, M. Sekiya, A. Yumoto, J. Yamada, T. Hirano, Y. Iwase, T. Yamada, T. Ishibashi, T. Mori, M. Asano, S. Tamura, and T. Urabe, “24.4L: Late-News Paper: A 13.0-inch AM-OLED Display with Top Emitting Structure and Adaptive Current Mode Programmed Pixel Circuit (TAC),” SID Symposium Digest of Technical Papers, vol. 32, no. 1, p. 384, 2001. [101] S. Ashtiani, P. Servati, D. Striakhilev, and A. Nathan, “A 3-TFT Current-Programmed Pixel Circuit for AMOLEDs,” IEEE Transactions on Electron Devices, vol. 52, pp. 1514–1518, jul 2005. [102] K. Sakariya, P. Servati, and A. Nathan, “Stability Analysis of Current Programmed a-Si:H AMOLED Pixel Circuits,” IEEE Transactions on Electron Devices, vol. 51, pp. 2019–2025, dec 2004. [103] J. Yamashita, K. Uchino, T. Yamamoto, T. Sasaoka, and T. Urabe, “44.2: New Driving Method with Current Subtraction Pixel Circuit for AM-OLED Displays,” SID Symposium Digest of Technical Papers, vol. 36, no. 1, p. 1452, 2005. References 95 [104] G. Chaji and A. Nathan, “Low-Power Low-Cost Voltage-Programmed a-Si:H AMOLED Display for Portable Devices,” Journal of Display Technology, vol. 4, pp. 233–237, jun 2008. [105] S.-H. Jung, W.-J. Nam, and M.-k. Han, “P-104: A New Voltage Modulated AMOLED Pixel Design Compensating Threshold Voltage Variation of Poly-Si TFTs,” SID Sym- posium Digest of Technical Papers, vol. 33, no. 1, p. 622, 2002. [106] Joon-Chul Goh, Hoon-Ju Chung, Jin Jang, and Chul-Hi Han, “A new pixel circuit for active matrix organic light emitting diodes,” IEEE Electron Device Letters, vol. 23, pp. 544–546, sep 2002. [107] S.-M. Choi, O.-K. Kwon, and H.-K. Chung, “P-11: An Improved Voltage Programmed Pixel Structure for Large Size and High Resolution AM-OLED Displays,” SID Sym- posium Digest of Technical Papers, vol. 35, no. 1, p. 260, 2004. [108] Jae-Hoon Lee, Ji-Hoon Kim, and Min-Koo Han, “A new a-Si:H TFT pixel circuit compensating the threshold voltage shift of a-Si:H TFT and OLED for active matrix OLED,” IEEE Electron Device Letters, vol. 26, pp. 897–899, dec 2005. [109] Y.-H. Tai, B.-T. Chen, Y.-J. Kuo, C.-C. Tsai, K.-Y. Chiang, Y.-J. Wei, and H.-C. Cheng, “A New Pixel Circuit for Driving Organic Light-Emitting Diode With Low Temperature Polycrystalline Silicon Thin-Film Transistors,” Journal of Display Technology, vol. 1, pp. 100–104, sep 2005. [110] Y. Matsueda, R. Kakkad, Y. S. Park, H. H. Yoon, W. P. Lee, J. B. Koo, and H. K. Chung, “35.1: 2.5-in. AMOLED with Integrated 6-Bit Gamma Compensated Digital Data Driver,” SID Symposium Digest of Technical Papers, vol. 35, no. 1, p. 1116, 2004. [111] C.-L. Lin and T.-T. Tsai, “A Novel Voltage Driving Method Using 3-TFT Pixel Circuit for AMOLED,” IEEE Electron Device Letters, vol. 28, pp. 489–491, jun 2007. [112] J. L. Sanford and F. R. Libsch, “4.2: TFT AMOLED Pixel Circuits and Driving Methods,” SID Symposium Digest of Technical Papers, vol. 34, no. 1, p. 10, 2003. [113] M. Bagheri, S. Ashtiani, and A. Nathan, “Fast voltage-programmed pixel architecture for AMOLED displays,” Journal of Display Technology, vol. 6, no. 5, pp. 191–195, 2010. [114] H. J. In and O. K. Kwon, “External compensation of non-uniform electrical character- istics of thin-film transistors and degradation of OLED devices in AMOLED displays,” IEEE Electron Device Letters, vol. 30, no. 4, pp. 377–379, 2009. [115] S. Jafarabadiashtiani, G. Chaji, S. Sambandan, D. Striakhilev, A. Nathan, and P. Ser- vati, “P-25: A New Driving Method for a-Si AMOLED Displays Based on Voltage Feedback,” SID Symposium Digest of Technical Papers, vol. 36, no. 1, p. 316, 2005. [116] J. H. Baek, M. Lee, J. H. Lee, H. S. Pae, C. J. Lee, J. T. Kim, C. S. Choi, H. K. Kim, T. J. Kim, and H. K. Chung, “A Current-Mode Display Driver IC Using Sample-and- Hold Scheme for QVGA Full-Color AMOLED Displays,” IEEE Journal of Solid-State Circuits, vol. 41, pp. 2974–2982, dec 2006. 96 References [117] S. J. Ashtiani and A. Nathan, “A Driving Scheme for Active-Matrix Organic Light- Emitting Diode Displays Based on Current Feedback,” Journal of Display Technology, vol. 5, pp. 257–264, jul 2009. [118] H.-J. In, B.-D. Choi, H.-K. Chung, and O.-K. Kwon, “Current-Sensing and Voltage- Feedback Driving Method for Large-Area High-Resolution Active Matrix Organic Light Emitting Diodes,” Japanese Journal of Applied Physics, vol. 45, pp. 4396–4401, may 2006. [119] H.-J. In, P.-S. Kwag, J.-S. Kang, O.-K. Kwon, and H.-K. Chung, “Voltage- programming method with transimpedance-feedback technique for threshold voltage and mobility compensations in large-area high-resolution AMOLED displays,” Jour- nal of the Society for Information Display, vol. 14, no. 8, p. 665, 2006. [120] R. Itou, M. Kayama, and T. Shima, “Some analog building blocks for TFT circuits,” in Proceedings of the 44th IEEE 2001 Midwest Symposium on Circuits and Systems. MWSCAS 2001 (Cat. No.01CH37257), vol. 1, pp. 417–420, IEEE, 2001. [121] P. R. Gray, P. Hurst, R. G. Meyer, and S. Lewis, Analysis and design of analog integrated circuits. Wiley, 2001. [122] R. Gregorian and G. C. Temes, Analog MOS Integrated Circuits for Signal Processing. New York: Wiley-Interscience, 1 ed., 1986. [123] Y. Tsividis, D. Fraser, and J. Dziak, “A process-insensitive high-performance NMOS operational amplifier,” IEEE Journal of Solid-State Circuits, vol. 15, pp. 921–928, dec 1980. [124] D. Senderowicz, D. Hodges, and P. Gray, “High-performance NMOS operational amplifier,” IEEE Journal of Solid-State Circuits, vol. 13, pp. 760–766, dec 1978. [125] M. J. Seok, M. H. Choi, M. Mativenga, D. Geng, D. Y. Kim, and J. Jang, “A Full- Swing a-IGZO TFT-Based Inverter With a Top-Gate-Bias-Induced Depletion Load,” IEEE Electron Device Letters, vol. 32, pp. 1089–1091, aug 2011. [126] B. Kim, C.-I. Ryoo, S.-J. Kim, J.-U. Bae, H.-S. Seo, C.-D. Kim, and M.-K. Han, “New Depletion-Mode IGZO TFT Shift Register,” IEEE Electron Device Letters, vol. 32, pp. 158–160, feb 2011. [127] D. Heineck, B. McFarlane, and J. Wager, “Zinc Tin Oxide Thin-Film-Transistor Enhancement/Depletion Inverter,” IEEE Electron Device Letters, vol. 30, pp. 514–516, may 2009. [128] H. Ma, R. W. R. Wallbank, R. Chaji, J. Li, Y. Suzuki, C. Jiggins, and A. Nathan, “An impedance-based integrated biosensor for suspended DNA characterization,” Scientific Reports, vol. 3, p. 2730, 2013. [129] K. Takechi, S. Iwamatsu, S. Konno, T. Yahagi, Y. Abe, M. Katoh, and H. Tanabe, “Demonstration of Detecting Small pH Changes Using High-Sensitivity Amorphous InGaZnO 4 Thin-Film Transistor pH Sensor System,” IEEE Transactions on Electron Devices, vol. 64, pp. 638–641, feb 2017. References 97 [130] M. Barron, “Low level currents in insulated gate field effect transistors,” Solid-State Electronics, vol. 15, pp. 293–302, 1972. [131] R. M. Swanson and J. D. Meindl, “Ion-implanted complementary mos transistors in low-voltage circuits,” IEEE Journal of Solid-State Circuits, vol. 7, no. 2, pp. 146–153, 1972. [132] R. Barker, “Small-signal subthreshold model for i.g.f.e.t.s,” Electronics Letters, vol. 12, p. 260, may 1976. [133] E. Fortunato, P. Barquinha, and R. Martins, “Oxide semiconductor thin-film transistors: A review of recent advances,” 2012. [134] T. Sekitani, U. Zschieschang, H. Klauk, and T. Someya, “Flexible organic transistors and circuits with extreme bending stability.,” Nature materials, vol. 9, no. 12, pp. 1015– 1022, 2010. [135] S. M. Sze and K. K. Ng, Physics of semiconductor devices. 2006. [136] Y. Nishi, “Insulated gate field effect transistor and its manufacturing method,” Japan Patent, vol. 587, no. 527, pp. 162–165, 1970. [137] M. P. Lefselter and S. M. Sze, “SB-IGFET: An Insulated-Gate Field-Effect Transistor Using Schottky Barrier Contacts for Source and Drain,” Proceedings of the IEEE, vol. 56, no. 8, pp. 1400–1402, 1968. [138] J. M. Shannon and F. Balon, “Source-gated thin-film transistors,” Solid-State Electron- ics, vol. 52, no. 3, pp. 449–454, 2008. [139] D. Sarkar, X. Xie, W. Liu, W. Cao, J. Kang, Y. Gong, S. Kraemer, P. M. Ajayan, and K. Banerjee, “A subthermionic tunnel field-effect transistor with an atomically thin channel,” Nature, vol. 526, no. 7571, pp. 91–95, 2015. [140] T. T. Trinh, K. Jang, V. A. Dao, and J. Yi, “Effect of high conductivity amorphous InGaZnO active layer on the field effect mobility improvement of thin film transistors,” Journal of Applied Physics, vol. 116, p. 214504, dec 2014. [141] Y. Xu, H. Sun, and Y.-Y. Noh, “Schottky Barrier in Organic Transistors,” IEEE Transactions on Electron Devices, vol. 64, pp. 1932–1943, may 2017. Appendix A Device Circuit Interaction A.1 Importance of DCI in TFTs Although the ever-evolving TFT technology continues to produce devices with improved performance, such as higher mobility, steeper subthreshold slope and lower VT [? ? ? ? ? ? ? ? ? ], circuit implementation is still somewhat constrained. This applies for most of the material families including metal-oxides, organics, and amorphous silicon (a-Si:H) although much less so with low temperature poly-silicon (LTPS). Here, a key design consideration is the device-circuit interaction (DCI), which has to be accounted for when circuits are designed with devices of poor performance and high degree of non-ideality [? ? ? ? ? ? ? ? ? ? ] as compared to the CMOS counterpart. This is particularly true when the intrinsic performance of TFTs does not the meet the requirements of a desired application. As shown in Fig. A.1, if the performance of the desired application is much lower than the maximum achievable intrinsic performance of the TFT, it is then possible to design the circuit independently without considering device non-idealities. For example, when the error in the TFT’s output current created by VT shift is much lower than the required accuracy, the VT shift problem is not of concern. However, when the performance requirement needs to be higher than the intrinsic performance, the designer should seek a compensation solution based on DCI or wait for improvements in the technology. We will discuss the intrinsic performance of TFTs in Section A.2 along with compensation methods in Section A.3. Another aspect of DCI stems from the material and processing attributes of the TFT which usually come with specific, and often self-limiting, properties. For example, in analogue front-end and digital designs, alternative circuit architectures are needed to match the properties of the CMOS counterpart [? ? ? ? ? ]. This will be discussed in Section A.4 along with solutions to deal with, for example, light-induced non-ideality in oxide TFTs. 100 Device Circuit Interaction Fig. A.1 Illustration of DCI in relation to performance requirements of a desired application A.2 Impact of TFT Properties The TFT is the major building blocks of active thin film circuits and its properties determine circuit performance. In applications such as displays and analogue front-end circuits, the accuracy of the output signal strongly affects the quality of the displayed image without mura (luminance non-uniformity) or in processing analogue signal without significant error. The critical parameters for TFTs (e.g. mobility, Cox, VT , etc.) determine the performance of the circuit and are more often discussed when comparing TFT behaviour or modelling a single transistor’s terminal characteristics [? ? ? ? ? ? ? ? ]. However, other issues such as stability, temperature sensitivity and process variations can also limit the overall performance and may even affect the functionality of the circuit. There has been significant effort devoted to the study and modelling of bias induced VT -shift [? ? ? ] and VT -shift compensation in AMOLED pixel circuits [? ? ? ? ]. We will analyse the sensitivity of drain current on VT -stability and temperature and process variations with the aim of establishing guidelines on the level of accuracy that can be achieved without applying compensation methods (i.e. the intrinsic performance) as well as identify the parameters that contribute most to error and how this can be improved with processing. A.2.1 Temperature Dependence The parameters of the TFT are affected by material properties and on temperature, which in turn impacts the terminal current-voltage (I-V) behaviour. According to the I-V model of the A.2 Impact of TFT Properties 101 Fig. A.2 I-V characteristics of the examined TFT under different temperatures TFT: IDS ≈ 12+αp µ∗0 Qαp−1re f W L′ Cαpox (VGS−VT )αp+1 ≡ K× (VGS−VT )αp+1 (A.1) where, K ≡ 1 2+αp µ∗0 Qαpre f W L′ Cαpox (A.2) The variations here can be specified as the variation of three key parameters: K, VT and αp. By considering these parameters as functions of temperature, the derivative of IDS as a function of temperature can be expressed as follows: dIDS(T ) = (VGS−VT )αp+1∂K∂T dT −K(αp+1)(VGS−VT )αp ∂VT∂T dT +Kln(VGS−VT )(VGS−VT )αp+1∂αp∂T dT (A.3) Therefore the temperature sensitivity of the overall current can be separated into three parts, in which each part of the function is determined by the temperature sensitivity of K, VT or αp. The contribution of each parameter can then be calculated through extraction of the temperature sensitivity of the three parameters. 102 Device Circuit Interaction Consider the transfer characteristic for an indium-gallium-zinc-oxide (IGZO) TFT, mea- sured every 10°C from 40°C to 80°C, shown in Fig. A.2. The results show that the overall current would increase when temperature increases. To further investigate the degree of influence of the three parameters, their values have been extracted from the measured transfer characteristics according to Eq. (A.1). Here, we extract the threshold voltage (VT ) indepen- dently from a multi-derivative method [? ], and then use it to calibrate the gate voltage as VGS−VT . With this, I-V data is plotted in log-log plot. In this plot, all the data turns into a linear behaviour, where the intercept on the y-axis, log(IDS), is log(K), with slope αp. From this, we will get K and αp independently. The results are shown in Fig. A.3. As can be seen, all three parameters are approximately linearly related to 1/kT in the temperature range considered. Therefore, through a linear fitting of the parameters, we get the empirical models of the parameters with the following relations: αp(T ) = α0+ Eα kT (A.4) K(T ) = K0exp ( −EK kT ) (A.5) VT (T ) =VT0exp ( −EV T kT ) (A.6) Combining Eq. (A.4) (A.5) (A.6) and (A.3), the current sensitivity can be derived as: dIDS(T ) dT = (VGS−VT )αp+1K0 EKkT 2 exp ( −EK kT ) −K(αp+1)(VGS−VT )αpVT0 EV T kT 2 exp ( −EV T kT ) −Kln(VGS−VT )(VGS−VT )αp+1 EαkT 2 (A.7) Note that the three terms at the RHS of Eq. (A.7) define the contribution of K, VT and αp, respectively. To understand how much the overall current is affected by ambient temperature and the associated contribution of the different parameters, the normalized temperature sensitivity is shown in Fig. A.4. As can be seen, the overall temperature sensitivity drops with increase in VGS due to the fact that the contribution of VT drops while VGS increases. This tendency starts to saturate when VGS increases to 4V, when the contribution of K becomes dominant. This analysis suggests that TFTs can be very unstable when biased at a voltage near VT . Although higher stability can be achieved through intentionally biasing the transistor at higher voltage A.2 Impact of TFT Properties 103 Fig. A.3 Extracted values of K, αp and VT at the different temperatures 104 Device Circuit Interaction Fig. A.4 Normalized temperature sensitivity of current and the contribution of different parameters at 313K levels, the maximum achievable level will be determined by the temperature sensitivity of K. As the temperature dependence of αp has a negative contribution to current with respect to temperature and its contribution increases at higher bias, the temperature dependence of K can be compensated by αp resulting in decreased sensitivity. However, a higher bias level implies increased power consumption of the circuit. Therefore, an appropriate bias point should be chosen for enough stability with acceptable power consumption. A.2.2 Geometric Dependence Apart from time- or temperature-dependent variations in device parameters, processing- induced spatial variations should be considered especially in pixelated arrays or analogue circuit applications. These variations would cause pixel non-uniformity in displays or imagers and create error or undesired behaviour in analogue circuit design. Note that the non-uniformity can be global (i.e. between panels) or local (i.e. between transistors). The latter is harder to deal with especially when transistor matching is of concern (such as in A.2 Impact of TFT Properties 105 differential pairs or current mirrors). This chapter will focus on the local non-uniformity and discuss the contribution of different parameters in creating the current mismatch. Analysis of geometric dependence can follow a similar route as with temperature depen- dence. As different parameters would follow a certain probability distribution, the overall current will be determined by the randomness of all three parameters according to Eq. (A.1). As variations are usually smaller than the respective mean values, the mismatch in IDS can be expressed as a first order approximation: ∆IDS = (VGS−VT 0)αp0+1∆K −K0(αp0+1)(VGS−VT 0)αp0∆VT +K0ln(VGS−VT 0)(VGS−VT 0)αp0+1∆αp (A.8) Here, K0, VT 0 and αp0 are the mean values of the parameters, and ∆K, ∆VT and ∆αp their respective variations. Assuming K, αp and VT are independent variables and that all follow a normal distribution, the variance of IDS can be expressed as: σ2IDS = (VGS−VT 0)2(αp0+1)σ2K +K20 (αp0+1) 2(VGS−VT 0)2αp0σ2VT +K20 [ln(VGS−VT 0)]2(VGS−VT 0)2(αp0+1)σ2αp (A.9) where σK , σVT and σαp are the standard deviation of K, αp and VT , respectively. As expected, the standard deviation of the overall current is determined by the deviation of all three parameters. To analyze the variation sensitivity of the overall current, we need statistical data for all three parameters. Here, we acquired statistical data of the transfer characteristics by measuring a 1080×1920 RGBW OLED display panel. By using the pixel circuit described in [? ], which will be reviewed in the next section, we could extract the I-V characteristics of the driver TFTs within the panel and extract statistical data for the three parameters. Here, the TFTs for driving the green OLED pixels are used as shown in Fig. A.5. The probability density functions are fitted to a normal distribution using MATLAB. With the fitted mean values and standard deviations, we calculate the relative contributions of each parameter to the current variance using Eq. (A.9). The standard deviation of IDS and the contribution of each parameter are shown in Fig. A.6. We see a similar overall curve in the sense that the sensitivity is higher when biased near VT . However, unlike the temperature dependence where sensitivity drops with increasing bias, the dependence on geometric shows a minimum at around VGS=6V. This is due to the decreasing contribution of VT and the increasing contribution of αp. Therefore, 106 Device Circuit Interaction (a) Fitted probability distribution and histogram of K (b) Fitted probability distribution and histogram of VT (c) Fitted probability distribution and histogram of αp Fig. A.5 Fitted probability distribution and histogram of the three key parameters. The data was extracted at room temperature from a 1080*1920 RGBW AMOLED panel with pixel circuits as described in Fig. A.7. A.3 Compensation Methods in Circuits & Applications 107 Fig. A.6 Normalized standard deviation of current and the contribution of different parame- ters(K, VT and α) analogue designers can intentionally choose a bias point close to the minimum point for the output transistor, when designing circuits to reduce the output current sensitivity to process variations. The analysis of temperature and geometric dependence here is done using a generic approach since the current-voltage behaviour is estimated by three key parameters, namely K, VT and αp. This is adaptable to most TFT types because of the similar working principle albeit with different parameter values. Therefore, the methodologies and derivations presented here are generic and empirically approached for applicability to other material families including OTFTs and related material families. A.3 Compensation Methods in Circuits & Applications . The intrinsic performance of TFTs does not always meet the requirements of specific applications. For example, the threshold voltage shift in TFTs creates visible shadows or ghosting in displays or imagers after extended operation. The resulting non-uniformity 108 Device Circuit Interaction creates mismatch in output characteristics especially in matrix architectures (displays or sensors). These kinds of defects have proven difficult to improve by processing, and thus need to be compensated through circuit solutions. Here, we will discuss VT shift and non- uniformity compensation methods such as on-pixel programming and by off-pixel feedback. The compensation methods can also be extended to other circuit applications. A.3.1 On-Pixel Programming The most common way of compensating TFT defects in active matrix arrays is through on-pixel compensation. The methods were first developed for a-Si:H TFTs in AMOLEDs as this family of TFTs have severe VT shift under positive bias, which leads to big errors when supplying current to the OLED [? ]. VT compensation techniques can be categorized into two kinds: current programming [? ? ? ? ? ? ? ? ] and voltage programming [? ? ? ? ? ? ? ? ]. The basic working principles of these have been reviewed in [? ? ]. Recent progress in voltage programming has been reported in [? ? ]. Here, the working principle is slightly different from the original idea in that the technique does not capture the cut-off point of a diode-connected transistor. Instead, it uses a TFT to discharge a pre-charged capacitor with fixed gate bias in a fixed time period to capture the TFT’s property (i.e. VT ). This technique can provide faster VT extraction and is good particularly when the speed of the circuit is of concern. A.3.2 Off-Pixel Feedback Another way of defect compensation is based on off pixel feedback. Since in display applications the driving period of each pixel can be separated into several phases, it is possible to use part of the driving sequence to extract all of the defect and aging data present in the pixel and then drive with revised extracted parameters as feedback [? ? ? ? ? ? ? ]. A pixel structure for defect extraction reported in [? ] is shown in Fig. A.7. The pixel circuit shows a similar structure as the simple 2T1C structure – the only difference being addition of a monitor line to monitor the TFT and OLED characteristics and extract their defect and aging status. The driving sequence for this pixel circuit is similar to the 2T1C except for the addition of a defect extraction phase. Defect extraction starts after the SEL signal selects the pixel and before writing data to it. The extraction phase consists of two parts – the driver TFT and the OLED, respectively (shown in Fig. A.8). Fig. A.8a shows an equivalent circuit of the extraction phase for the OLED. In this phase the gate voltage of the driver TFT is set to ground level to turn it off and the monitor line is set to a higher voltage of VOLED. The current flowing through the OLED can then be A.3 Compensation Methods in Circuits & Applications 109 Fig. A.7 AMOLED pixel structure for off-pixel defect extraction and feedback captured by measuring the current flowing into the monitor line. As the I-V characteristic of the OLED can be a signature of its efficiency, the aging of OLED can then be captured from pre-acquired data for this type of OLED. In Fig. A.8b, the gate voltage of the driver TFT is set to a higher level of VP. The TFT is then turned on and part of the drain current flows to the monitor line. Combining the extracted I-V characteristic data of the OLED and the voltage and current measurements in this phase, the I-V characteristic of the driver TFT can be extracted. After capturing the defect and aging data for both the driver TFT and OLED, the data voltage for the desired luminance can be calculated and applied to the pixel by external circuitry. The aging and defect status of TFTs and OLEDs extracted through the monitor lines are shown in Fig. A.9(a) and Fig. A.9(b), respectively. The sharp hazards that happen randomly among the panel show the fabrication defects of the pixels. And the patterns in red, yellow and green colour express the aging of each device. As the center of the displayed white square has higher temperature due to self-heating of the surrounding pixels, the aging of these pixels will be faster compared to other pixels. The display panel using this method has the ability of compensating all kinds of defects that can be extracted through the monitor line. The compensation results are shown in Fig. A.10 depicting the defect status in Fig. 10(a)&(b) and temperature compensation in Fig. 10(c)&(d). 110 Device Circuit Interaction (a) OLED detection phase (b) TFT detection phase Fig. A.8 Equivalent circuit for the defect extraction phase of OLED and TFT A.3 Compensation Methods in Circuits & Applications 111 Fig. A.9 Extracted aging parameters for (a) TFTs and (b) OLEDs after continuously display- ing a checker board (W: displayed with white squares, B: displayed with black squares) Fig. A.10 Extracted aging parameters for (a) TFTs and (b) OLEDs after continuously displaying a checker board (W: displayed with white squares, B: displayed with black squares) 112 Device Circuit Interaction This method provides a generic solution for TFT compensation and will be particularly useful in compensation of mechanically-induced (reversible) defects or aging in flexible displays. As measurement of the TFTs can be done in a monitoring phase, the obtained parameters can then be used to bias the TFTs to have an ideal overall performance. The geometric dependence of transistors can also be extracted from this pixel, as during the sequence described in Fig. A.8a & A.8b one can also apply voltage sweep to obtain the measured I-V characteristics for both TFT and OLED devices. In summary, defects of TFTs can be compensated through separating the working se- quence into several phases and by adding a compensation phase to extract and compensate non-idealities. To apply this method to other analogue circuits, it is possible to intentionally separate the working sequence and apply similar methods. Here, switch-capacitor circuits can be a promising candidate. An example of this applied to analogue building blocks was reported in [? ]. A.4 Specific Device Properties and Alternate Circuit Ar- chitecture Devices processed using different thin film technologies usually have a specific property, which can sometimes be self-limiting. This requires the use of alternate circuit architectures. For example, most TFT technologies lack complementarity, thus the circuits have to be designed using mono-type devices, which means the designs cannot benefit from the well- established CMOS architectures. Specifically, the load of an analogue amplifier needs to be redesigned to achieve high gain as the complementary load is not applicable. Another example is persistence photoconductivity in oxide TFTs, which, while ideal for image capture, requires a sharp gate a pulse to reset for high frame rates. We will discuss these examples in the following. A.4.1 Analogue Gain Stage with Mono-type TFTs CMOS gain stages benefit from the complementary structure [? ? ? ]. For example, in the case of NMOS as the input stage and PMOS as the load, the gain of the stage can be boost up to the order of magnitude of gmro. The use of PMOS load provides large enough bias current with small voltage and, at the same time, big small signal resistance. These requirements are hard to be achieved with only one type of transistors. To simplify the problem, we consider a single common source amplifier stage as shown in Fig. A.11(a). Here, the load is considered as a two terminal device. Assuming the bias A.4 Specific Device Properties and Alternate Circuit Architecture 113 Fig. A.11 (a) Common-source gain stage of an amplifier (b) sketch of different type of loads passing through the same bias point conditions of the driver TFT is fixed (to maintain gm for fair comparison between different loads), the load of this gain stage would have fixed bias current. Here, we consider a fixed bias point of the load for comparison, i.e. the I-V characteristic of the load should pass a fixed point in the I-V plot (Fig. A.11(b)). As the gain stage needs a higher small signal resistance of the load to produce high gain (i.e. derivative at the bias point should be close to zero), the current is better a concave function of voltage which means p-type load is more beneficial. Note that for other type of loads, the same small signal resistance can only be achieved by moving the bias point to the right hand side, i.e. increase the voltage bias of the load. However, a higher voltage bias would yield higher power consumption. From this standpoint, we conclude that for high gain, a p-type load with fixed gate bias (concave function) is more beneficial than a resistive load (linear function) and, subsequently, a diode or diode-connected TFT load (convex function) as it yields higher gain at the same power consumption. The most familiar concave function in TFT behaviours is the output characteristic (ID− VDS characteristic with fixed VGS). However, the design is limited by the connection of n-type devices as the source terminal of the load TFT is connected to the output node of the gain stage. Thus, the gate terminal of the TFT cannot be chosen as a fixed level but should follow the change of the source terminal. 114 Device Circuit Interaction Fig. A.12 Gain stage of an amplifier with (a) feedback load (b) boost strap load One straight forward solution is to use a depletion mode load and short circuit the load transistor’s source and gate terminals. (Depletion mode transistor can work in saturation regime with zero VGS.) Amplifiers have been designed around 1980s to achieve high gain with depletion and enhancement NMOS transistors [? ? ? ]. However, in TFT circuit area, this approach is limited by process complexity. By far, only digital circuits have been fabricated out of depletion load [? ? ? ]. In order to use only enhancement mode TFTs while still obtaining high enough gain, it is also possible to use feedback loop to maintain a fixed VGS for high small signal resistance of the load. One approach is reported in [? ? ]. The circuit reported is shown in Fig. A.12(a). Here, an analogue adder is designed for the DC bias of the load transistor. The feedback circuit, in effect, adds a bias voltage to the source terminal of the load transistor and applies the resulting voltage to the gate terminal. Here, the feedback voltage at the gate terminal of the load TFT can be calculated as: VF =VH −VB+VO (A.10) Here, VH and VB are external bias voltages and VO is the output of the gain stage (which is also the source terminal voltage of load). As the feedback voltage contains the output, this topology has a positive feedback. Another approach using positive feedback has been reported by H.Marien et al [? ]. The load of their gain stage is designed with boost strap structure. Other than maintaining the A.4 Specific Device Properties and Alternate Circuit Architecture 115 VGS level of the load TFT, the structure successfully separates the DC bias and the small signal resistance and as a result obtains a higher gain for small signals at a higher frequency. The circuit needs a large capacitor for the load (a high pass filter) to reduce its lower 3dB frequency close to DC range. This also makes the circuit hard to work at very low frequencies, especially when the original signal is close to DC. An NMOS version of the circuit has been shown in Fig. A.12(b). Other approaches regarding positive feedback are also reported in [? ]. To summarize, positive feedback is used for gain-enhancement in single type TFT amplifiers. The major problem of this kind of approach is that positive feedback sacrifices the phase margin (PM) of the amplifier and would potentially cause instability. Although a cascode technique can be chosen to enhance the gain without sacrificing PM, it is not preferred in TFTs since a much higher supply voltage is needed because of the high threshold voltage of TFTs. A.4.2 Persistent Photoconductivity This can arise depending on the channel composition in oxide transistors when the transistor is under exposure to ambient light or when subject to negative bias illumination stress (NBIS) [? ][21]. As the drain current of the TFT increases after exposure to ambient light, it is possible to use this characteristic for photo-sensing applications. In this application, the variation of current is not a defect that needs to be compensated. However, to use this property, the slow recovery process makes it hard to capture the changing image at high frame rate or to restore the TFT’s initial state. Fig. A.13(a) depicts the drain current of an IGZO TFT under periodic luminance and darkness. The results show the drain current recovering only slowly even in a completely dark environment. In order to remove the persistent photoconductivity and recover the TFT to its original state, it has been found that by applying a positive pulse to the gate of the TFT, the PPC can be eliminated very quickly. The results in Fig. A.13(b) show that fast PPC removal is possible after the gate pulse technique. The phenomenon of PPC is explained by the band diagram shown in Fig. A.14. As the ambient light excites the electrons in defect states to the conduction band, ionized oxygen defects are created. The excited electrons effectively increase the conductivity of the device as the electron concentration is increased. In order to accelerate the recovery process, the recombination of photo-induced electrons and ionized oxygen defects VO2+ needs to be accelerated. By applying a positive voltage pulse to the gate of the transistor, the electrons are swept away from the conduction band to recombine with the ionized oxygen defects thus removing the PPC. 116 Device Circuit Interaction Fig. A.13 Band diagram to explain the effect of positive gate pulse on recovery. PPC removal is a great example of device circuit interaction as the defect (i.e. light induced instability) of a TFT can actually be utilized to sense luminance signal, i.e. as a photo sensor. Indeed DCI should be considered not only to enhance circuit performance and/or overcome difficulties in the design but also to utilize specific properties stemming from the operating environment. A.5 Conclusion This chapter reviewed and analysed the intrinsic parameters of TFTs and proposed ways of analysing the maximum achievable current accuracy of TFTs and to help determine when compensation becomes mandatory to enhance reliability and combat ageing. We presented techniques for extraction of defects and aging in devices using closed-loop feedback techniques and discussed their extension to other applications. In summary, device circuit interactions are crucial when designing high performance circuits and systems to either utilize or minimize the impact of intrinsic adversaties associated with low temperature thin film technology. Consideration of device circuit interactions in design of TFT systems is even more compelling than the case of CMOS technology because of the wide range of materials imperfections, which give rise to device instability and large area processing- induced non-uniformity. Of specific importance to mechanically flexible systems is the impact of bending-induced (reversible) defects and associated aging, which makes compensation A.5 Conclusion 117 (a) Observation of persistent photoconductivity (PPC) as a slow recovery (b) Removal of PPC with a positive gate pulse to get fast recovery Fig. A.14 Persistent photoconductivity (PPC) and its removal 118 Device Circuit Interaction even more compelling. The techniques presented here can extend the current application of TFTs from active matrix pixelated arrays to newly-emerging application area that require TFT operation in analogue operation mode. Appendix B Subthreshold Operation for Schottky Barrier TFTs B.1 Introduction Recent research has shown that by biasing IGZO TFTs in deep subthreshold region could improve the small signal gain and significantly reduce the power consumption [? ]. This is specifically useful in low power applications such as sensor networks, bio-medical sensing etc [? ? ]. While subthreshold operation in Silicon based CMOS devices had been intensively researched in 1970s with great success in Swiss watch industry [? ? ? ? ], the same has not been done for TFTs. Due to the low effective carrier mobility of TFTs (perhaps with polysilicon as an ex- ception), TFT devices are generally working in a lower current level [? ? ]. Although this could limit the speed of the device, TFTs can be naturally suitable for low speed and low power applications. Biasing TFTs in subthreshold region could further reduce the power consumption to sub-nano-watt. This is particularly appealing as battery-less operation could be possible with this level of power consumption. However, as the subthreshold region is generally very narrow and current level is very sensitive to bias [? ]. If a TFT has a steeper subthreshold slope (SS), a sensitivity of current variation is higher. The circuit design for this kind of devices could be challenging although a steep SS provides a high intrinsic gain. Here, a high intrinsic gain is an important property for circuit design, so many approaches to get it have been introduced for the above-threshold operation [? ? ? ? ]. For the subthreshold operation, it is known than a TFT with a Schottky contact source/drain have a steeper SS 120 Subthreshold Operation for Schottky Barrier TFTs compared to an Ohmic contact TFT [? ? ? ], thus a Schottky contact TFT provides a higher intrinsic gain within the subthreshold regime with a difficulty in a circuit design. In this chapter, we will analyze several figures of merit for deep-subthreshold operating TFTs in analog circuit design with the aim of providing a guideline of analog circuits design capability of TFT devices using general architectures. B.2 Subthreshold Model for IGZO TFTs B.2.1 DC Model The transfer curve of SB-TFTs in deep-subthreshold is reported to be linear in log-scale which is similar to CMOS devices. Here we convert the model reported in [? ] to a more macroscopic form to help analyze the performance. The equation for drain voltage (VDS) > saturation voltage (vsat) is derived as follows [? ? ]: Ids = I′0 exp ( VGS−VT SS/ ln10 )( 1+ VDS VA ) (B.1) where I′0 is the effective subthreshold reference current at VT normalized by W and SS is the subthreshold slope. VT is the threshold voltage and VA is the effective early voltage. The effective values can be derived as I′0 ≡ AJJ0 exp ( VT −Vre f SS/ ln10 ) (B.2) where J0 is the reference current density at Vre f , AJ is the junction area of the Schottky contact at the source and drain. In a Schottky contact TFT, AJ is proportional to the channel width (W) [? ]. VA ≡ n · vth exp ( vsat nvth ) (B.3) where n is the ideality factor of the junction and vth is the thermal voltage. With the above equations we can easily connect the DC parameters to several important small signal parameters such as gm and ro. B.2.2 Small Signal Model Although in previous publications [17], [18] we analysed the small signal mode for TFTs and figured out that contact resistance can cause big errors using CMOS small signal models, B.3 Figures of Merit 121 small signal behaviour in subthreshold region can still be correctly modelled by CMOS model. This is due to the fact that channel capacitance in subthreshold is very small and normally negligible compared with overlap capacitance in TFTs. Here we consider Cgs and Cgd are mainly the overlap capacitance at source and drain side [17]. The tranconductance and output resistance can be calculated as: gm = ln10SS I ′ 0 exp ( VGS−VT SS/ ln10 )( 1+ VDSVA ) = ln10SS IDS (B.4) ro = IDS VA (B.5) These equations take exactly the same form as the CMOS counterpart. B.3 Figures of Merit In this section, we consider several figure of merit for TFTs with emphasize in the subthresh- old region. The model used for the simulation is based on the [? ] and parameter values are extracted using the same samples. The W/L of the TFT under test is 50µm/20µm. B.3.1 Intrinsic Gain Intrinsic gain is an important figure of merit for analog amplifier as it reflects the highest achievable single stage gain for an amplifier. The simulation based on the model developed earlier is shown in Fig. B.1. As seen, the intrinsic gain in subthreshold region is related with subthreshold slope and the ideality factor n of the Schottky junction. The expression of the intrinsic gain can be approximated as: Ai = gmro = ln10 VA ·SS (B.6) This suggests that intrinsic gain is reverse proportional to the SS of the TFT. By pushing the SS to its theoretical limit (60mV), one could get over 1000 intrinsic gain with even less ideal Schottky-Barrier at source-semiconductor contact (smaller n). The expression also suggests that the value of intrinsic gain is rather independent on the bias of the transistor as long as it is working in subthreshold region. Ideality factor n contributes to the intrinsic gain of the transistor because it controls the effective Early voltage (VA) and thus influence the output resistance of the TFT. 122 Subthreshold Operation for Schottky Barrier TFTs Fig. B.1 Simulation result for the intrinsic gain (Ai) of the Schottky-Barrier IGZO TFT (SB-TFT). The parameter n is the ideality factor of the source-semiconductor junction. B.3.2 Transconductance Efficiency Transconductance efficiency (gm/IDS) is another important figure of merit for electron devices. It represents the efficiency of converting the bias current into an equivalent transconductance. For analogue TFT circuits, as the voltage bias is normally limited by the driving circuitry and the application, reducing the current bias while maintaining a high gm also means reducing the power consumption while keeping the circuit performance. As can be seen from Fig. B.2, the highest gm/IDS can be obtained only at deep- sub- threshold region for a TFT before it enters a transition region to above-threshold. This value remains a constant at deep-subthreshold due to the exponential nature of the TFT subthreshold behaviour. The simulation result also suggests that this value would increase if the TFT has a steeper subthreshold slope. The expression of the gm/IDS value can be estimated as: gm IDS = ln10 SS (B.7) For above-threshold region, the gm/IDS value drops as the TFT enters above-threshold region, where the expression is: B.3 Figures of Merit 123 Fig. B.2 gm/IDS from deep subthreshold to above threshold region. The value reaches maximum at deep subthreshold and increases with steeper SS. gm IDS = 2+α VGS−VT (B.8) where α is the coefficient on the power law of TFT model, α=0 for the case of MOSFET [17]–[19]. Therefore, biasing a TFT at a high gate voltage would result in a less efficient gm conversion and thus higher power consumption. B.3.3 Cut-off frequency Cut-off frequency ( fT ) is also considered as the current gain bandwidth product as the current gain at fT is equal to 1 [17]. This value will limit the actual gain bandwidth product of amplifiers designed by the device. The simulation result of fT is shown in Fig. B.3. Here fT is small as the overlap capacitance of the TFT under test is quite big. The overlap length here is 50µm, which is even bigger than the channel length of the device. Practically, this value can be reduced to a few µm or even lower with self-aligned process. 124 Subthreshold Operation for Schottky Barrier TFTs Fig. B.3 Cut-off frequency vs. voltage bias for different SS. Fig. B.3 suggests that this product rolls off pretty quickly in subthreshold region. The expression of fT can be derived as follows: fT = gm 2πCov = IDS · ln10 SS ·2πCov (B.9) The expression shows that the value of fT is proportional to IDS, thus the quick roll of at subthreshold is explained by the current roll off in subthreshold region. As IDS and Cov are both proportional to the channel width of the TFT, fT would be independent of channel width. In addition, by reducing the overlap length, one could further reduce the overlap capacitance (Cov) to increase the frequency response. The expression also suggests that biasing the TFT in a deeper subthreshold region would at the same time pushing the frequency response lower proportionally. Therefore, it’s maybe desirable to have a less steep subthreshold slope in terms of speed when considering subthreshold operation. B.4 Sensitivity to Variations and Bias 125 Fig. B.4 Normalized sensitivity to VT shift for different SS B.4 Sensitivity to Variations and Bias B.4.1 Current Sensitivity to Variations Here, we assume the threshold voltage shift is also a uniform shift in subthreshold. A non-uniform shift in subthreshold would mean a change in subthreshold slope, which will be discussed later. Therefore, the VT shift at abovethreshold is the same as a reference voltage shift at subthreshold. As the reference voltage shift is equivalent to a bias voltage shift. This curve can also be used for testing the sensitivity of bias voltage inaccuracy. Fig. B.4 shows that at deep subthreshold region the normalized current sensitivity to VT shift is a constant. The sensitivity becomes lower when the TFT is approaching and enters above threshold region. As VT shift is equivalent to bias shift, the sensitivity expression can be the same as the gm/IDS curve (as gm = dIDS/dVGS is also a sensitivity to bias) just with an opposite sign. For deep-subthreshold: dIDS dVGS /IDS =− ln10SS (B.10) For above-threshold: dIDS dVGS /IDS =− 2+αVGS−VT (B.11) 126 Subthreshold Operation for Schottky Barrier TFTs Fig. B.5 Normalized sensitivity to SS variations for different SS mean values Now we assume the change in SS would not affect VT of the above threshold region. Therefore, the subthreshold current of different SS would meet at around VT . This yields the results shown in Fig. B.5, where the normalized current sensitivity would drop when the TFT is biased closer to VT . In above threshold region, the parameter SS do not contribute to the equation of IDS. The sensitivity is therefore zero. The curve also shows that a steeper SS would increase the current sensitivity. The expression for the current sensitivity in subthreshold is shown below: dIDS dSS /IDS =−(VGS−VT ) · ln10SS2 (B.12) The above equation suggests that the sensitivity is reverse proportional to SS2. Therefore, a steeper SS would yield a high current sensitivity on the variations of SS. In summary, although steeper SS would give rise to gm/IDS and potentially further reduce the power consumption, the sensitivity on process variations of SS and VT would increase, possibly narrowing the design window for analog circuits. B.4 Sensitivity to Variations and Bias 127 Fig. B.6 The useful bias range for a depletion load common-source amplifier. (a) conceptual figure of the load line and transfer curve of the amplifier, where green curve illustrates the correct bias condition and blue curves illustrates that the amplifying TFT is biased out of useful bias range (b) circuit schematic of the amplifier (c) simulated useful bias range with respect to SS B.4.2 Accuracy Requirement for Biasing Circuitry Here, we take a common source amplifier with depletion load as example (Fig. B.6(b)). The load line and bias conditions is shown in Fig. B.6(a). As can be seen, bias conditions should keep both transistors working in saturated region, i.e. the VGS for the green line in Fig. B.6(a). The blue lines show either of the two TFTs is working in a not saturated region. Therefore, to maintain a high gain, VDS for both TFTs should follow VDS > vsat . As the value of vsat for different VGS stays the same in subthreshold, the input voltage range can be calculated as: Vrange = 2 ·SS (VDD−2vsat) VA ln10 (B.13) 128 Subthreshold Operation for Schottky Barrier TFTs B.5 Conclusion We investigate several figures of merit of TFT’s subthreshold operation including Ai, gm/IDS, fT and the design sensitivity for biasing circuitry and device variations. The results of this chapter could benefit researchers working in device fabrication and circuit design while considering subthreshold operating TFTs specifically from a device performance perspective to a lower level analogue design. The results suggest that one could get high Ai, gm/IDS through biasing TFTs in subthreshold region at the cost of reduced fT and increased sensitivity to bias and process variations. Another interesting conclusion is that TFTs with less steep subthreshold slope can still benefit from a high Ai and gm/IDS, but with reduced sensitivity to variations and less strict requirement on accuracy of biasing circuitry.