See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/339787356 Simulation and Modeling of Tunnel Field Effect Transistor Thesis · November 2017 CITATIONS READS 0 1,747 2 authors: Prateek Jain Yogesh Singh Chauhan Indian Institute of Technology Kanpur Indian Institute of Technology Kanpur 3 PUBLICATIONS 23 CITATIONS 415 PUBLICATIONS 4,924 CITATIONS SEE PROFILE SEE PROFILE Some of the authors of this publication are also working on these related projects: Modeling and simulation of III-V and Ge transistors for logic and power applications View project Ferroelectric and Negative Capacitance Transistors - Modeling, Simulation and Circuit Design View project All content following this page was uploaded by Yogesh Singh Chauhan on 16 October 2020. The user has requested enhancement of the downloaded file. Simulation and Modeling of Tunnel Field Effect Transistor A Thesis Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy (Ph.D.) by Prateek Jain (12204070) to the Department of Electrical Engineering Indian Institute of Technology Kanpur November, 2017 ii iii iv SYNOPSIS Name of the Student: Prateek Jain Roll Number: 12204070 Degree for which Submitted: Ph.D. Department: Electrical Engineering Thesis Title: Simulation and Modeling of Tunnel Field Effect Transistor Names of the Thesis Supervisors: Dr. Yogesh Singh Chauhan and Dr. Amit Agarwal Month and Year of the Submission: November, 2017 The continuous scaling of transistors and the increase in transistor density, have led to immense increase in power density on chip. Supply voltage scaling reduces the dynamic power consumption, which is desirable for power constrained applications, but to achieve the same ON-current, threshold voltage Vth has to be scaled proportionately. The reduction in Vth causes exponential increase in OFF-current and static leakage power. This is because of subthreshold swing limit of 60 mV/decade at room temperature, due to the Boltzmann distribution of carriers. Therefore, it becomes imperative to explore new devices structures, which can operate at low voltage and consume less power. To overcome these issues, devices such as Tunnel Field Effect Transistor (TFET) have been explored, and also different architectures (Fully depleted silicon on insulator and FinFET) and different channel materials (MoS2 , III-V and Graphene) are attracting attention of the electronics community. In advanced technology nodes, fully and partially depleted silicon on insulator (FD-SOI and PD-SOI) and FinFET are some of the competitors. The devices, which are based on some other transport phenomena like interband tunneling and impact ionization are also being explored. TFET, which is based on interband tunneling phenomenon has attracted lot of attention of the electronics community, because of its excellent subthreshold slope and low off current. The main stumbling blocks with TFET are its low ON current and stringent fabrication steps, which restrict its usefulness for the mainstream device. Various optimization techniques have been reported to boost the performance of TFETs. In this work, we have investigated some of the optimization techniques. Further, first principle DFT v calculations are also performed to study the thickness dependent quantum confinement effects in Ge source lateral TFETs. We have also presented a surface potential based analytical model for TFET accounting the mobile charges in the channel. The thesis is organized as follows: In Chapter 1, we discuss the challenges involved in CMOS scaling and the available solutions for it. Tunnel field effect transistor is described in detail with the underlying physics and assumptions involved. The shortcomings of TFETs and the different proposed solutions to tackle them are also highlighted in this chapter. In Chapter 2, we have investigated “Dual Metal Gate (DMG)” technique in detail for the mono and hetero dielectric gate material with Si and relaxed SiGe sources. Dual metal gate technique allows to individually control the different portions of the transfer characteristics and this results in improved performance. The hetero dielectric at the gate reduces OFF current as compared to mono dielectric. Additionally, DMG with SiGe source offers the flexibility to modulate the band gap of SiGe at the source. Thus SiGe performance is found to be better than Si DMG-DGTFETs. In Chapter 3, we have examined the effect of compressive strain at the SiGe source and tensile strain in Si for Double Gate Tunnel Field Effect Transistor (DGTFET) with strained SiGe source. Sandwitched SiGe/Si/SiGe source is presented, which results in compressive strain at the source. The lattice constant of pseudomorphically grown SiGe on Si substrate is higher, as a result compressive strain is generated at the source. Further, at the vertical interface between SiGe/Si source, Si channel is tensile strained. The effect of pocket implant on the source for this structure is also discussed. The source extended structure is also examined to find the optimum source extension length. In Chapter 4, extended Ge source lateral TFET is investigated in detail. The direct and indirect valleys in Germanium (Ge) are separated by a very small offset, which opens up the prospect of direct tunneling in the Γ valley of extended Ge source TFET. We explore the impact of thickness scaling of extended Ge source lateral TFET on the band to band tunneling (BTBT) current. Ge source is extended inside the gate upto 2 nm to confine the tunneling in Ge only. We observe that as thickness is scaled, the band alignment at Si/Ge hetero junction changes significantly, and this results in increase of Ge to Si BTBT current. Based on density functional theory calculations, we first obtain the bandstructure parameters (bandgap, effective masses vi etc.) for the Ge and Si slabs of varying thicknesses, and these parameters are used to obtain the thickness dependent Kane’s BTBT tunneling parameters. Thereafter, Kane’s BTBT parameters of Si and Ge are used to obtain the scaled BTBT tunneling current. In Chapter 5, we have presented a comparative study of vertical and lateral tunneling in Ge source TFET as the source is scaled down from bulk to ultra-thin slabs of 4 nm thickness. We have theoretically calculated BTBT parameters for ultra-thin slab using band structure results from atomistic simulations, whereas for bulk we have considered bulk parameters. In our DGTFET structures, we have used an overlapped gate over Ge source to include vertical line tunneling along with lateral tunneling in the device. First, the impact of the gate overlap length scaling on DGTFET has been investigated at constant Ge source thickness of 8 nm. We have then analyzed the impact of scaling the Ge source thickness on DGTFET keeping the gate overlap length fixed. We have observed that Ge source thickness scaling till 8 nm has negligible effect on the ON current, but as the thickness is further scaled down, line tunneling disappears in the device. We also examine TFETs with ultra-thin Ge source, but with small gate-source overlap, to confine the tunneling in Ge only. In this configuration, lateral tunneling drives the drain current, leading to better sub-threshold slope and ON current. Further, we highlight the roles of enhanced subsurface BTBT current and gate screening in these ultra-thin lateral TFETs. In Chapter 6, we have developed a surface potential based analytical model for DGTFET for the drain current, terminal charges, and terminal capacitances. The model accounts for the effect of mobile charges in the channel and captures the device physics in depletion as well as in the strong inversion regime. The narrowing of tunnel barrier, in the presence of mobile charges in the channel, is incorporated by modeling of the inverse decay length, which is constant under channel depletion condition and bias dependent under inversion condition. To capture the ambipolar current behavior in the model, tunneling at the drain junction is also included. The proposed model is validated against TCAD simulation data and it shows close match with the simulation data. In Chapter 7, we summarize the research work presented in this thesis and also suggest the scope of future work in this area. Dedicated to My Grandparents Acknowledgement Above all, I thank my parents for their love and moral support. Living away from family is not easy, but they always make me feel like they are there for me. This would not have been possible without my parents. I sincerely thank my supervisor, Dr. Yogesh Singh Chauhan and Dr. Amit Agarwal for giving me the opportunity to work with them. Along with my academic research activities, they continually encouraged me to explore industrial experiences. All this is possible today, because they believed in me, more than I believed in myself. I would also like to thank my previous thesis supervisor Dr. Bahniman Ghosh, Dr. Dipanjan Basu, and Dr. Shyama Prasad Das for there guidance. I am deeply indebted to my teachers at IIT Kanpur, especially Prof. Aloke Dutta, Prof. B. Mazhari, Prof. S.S.K. Iyer and Prof. S. Qureshi for giving me deep insight and proper background during the course work. I would like to convey my special regards to Dr. Sarvesh Chauhan for sharing his memories of thesis writing, as it helps me a lot in improving my way of thesis writing. I am very thankful to my labmates and friends Chandan Yadav, Priyank Rastogi, Chetan Dabhi, Sudip Ghosh, Ravi Verma and Dinesh for useful discussions. Special thanks to Chetan Dhabi and Ravi Verma as their jokes were very helpful at the time of work pressure. I would like to thank Mr. Johnson Whiteford for his timely support in providing accessibility to various TCAD tools. I forever owe utmost gratitude to almighty GOD, whose blessings provide me faith and strength to carry on my work. viii Contents List of Figures xii 1 . . . . . . . . . . . 1 2 4 5 5 6 6 7 9 10 12 14 . . . . . . . . . . . . 25 25 27 29 30 31 33 34 35 36 37 39 39 Strain Engineering in SiGe source TFETs 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Device structure, fabrication, and simulation models. . . . . . . . . . . . . . . 3.2.1 Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 45 47 47 2 3 Introduction 1.1 CMOS Crisis . . . . . . . . . . . . . . . . . . 1.2 Alternative Solutions to Scale CMOS . . . . . 1.2.1 Alternative Channel Materials . . . . . 1.2.2 Alternative Device Architectures . . . . 1.2.3 Alternative Device Phenomena . . . . . 1.3 Tunnel Field Effect Transistor . . . . . . . . . 1.3.1 MOSFET vs TFET . . . . . . . . . . . 1.4 Operation of Tunnel Field Effect Transistor . . 1.4.1 Shortcomings of the TFETs . . . . . . 1.5 Circuits Performance Using InAs HomoTFETs 1.6 Thesis Goals and Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dual Metal - Double Gate Tunnel Field Effect Transistor 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 DMG-DGTFET: Schematic, fabrication and simulation. . . . . . . 2.3 DGTFET based on hetero dielectric gate material . . . . . . . . . . 2.4 Workfunction engineered DGTFET with HD gate . . . . . . . . . . 2.4.1 Workfunction engineered - Auxiliary Gate . . . . . . . . . . 2.4.2 Workfunction engineered - Tunnel Gate . . . . . . . . . . . 2.4.3 Output characteristics . . . . . . . . . . . . . . . . . . . . . 2.5 DMG-DGTFET with SiGe Source . . . . . . . . . . . . . . . . . . 2.5.1 Workfunction engineered DMG-DGTFET with SiGe source 2.5.2 Effect of gate dielectrics . . . . . . . . . . . . . . . . . . . 2.5.3 Effect of Ge mole fraction in hetero SiGe source . . . . . . 2.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . CONTENTS 3.3 3.4 3.5 3.6 3.7 x 3.2.2 Simulation models . . . . . . . . . . . . . . . . . . . . . . . . . . Strain Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.1 Compressive Strain in SiGe . . . . . . . . . . . . . . . . . . . . . 3.3.2 Tensile strain in Si . . . . . . . . . . . . . . . . . . . . . . . . . . BTBT Model Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . Field Induced Quantum Confinement (FIQC)- Back Envelope Calculations. Results and Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6.1 DGTFET with HD gate and c-SiGe source . . . . . . . . . . . . . 3.6.2 DGTFET with Pocket Implant . . . . . . . . . . . . . . . . . . . . 3.6.3 DGTFET with source extended structure . . . . . . . . . . . . . . 3.6.4 Channel length scaling . . . . . . . . . . . . . . . . . . . . . . . . Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 48 49 49 50 52 53 53 56 58 61 61 4 Band to Band Tunneling in Gamma Valley For Ge Source TFET: Thickness scaling 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Device Structure and Simulation Details . . . . . . . . . . . . . . . . . . . . . 4.2.1 Device Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.2 Atomistic Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.3 TCAD Simulation Models . . . . . . . . . . . . . . . . . . . . . . . . 4.3 BTBT Parameter and Current Calculations . . . . . . . . . . . . . . . . . . . . 4.3.1 Effective Masses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.2 Kane Parameters A and B . . . . . . . . . . . . . . . . . . . . . . . . 4.3.3 Effect of DOS in the effective electric field . . . . . . . . . . . . . . . 4.3.4 Effect of band alignment at Si/Ge junction . . . . . . . . . . . . . . . . 4.4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4.1 Ge to Si: Band alignment and BTBT . . . . . . . . . . . . . . . . . . . 4.4.2 Ge to Ge:Band alignment and BTBT . . . . . . . . . . . . . . . . . . . 4.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 66 67 67 68 69 70 70 70 71 72 73 74 76 78 5 Vertical and Lateral Tunneling in Ge Source TFET: Thickness Scaling 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Device Structure and Simulation Models . . . . . . . . . . . . . . . 5.3 BTBT Model Parameters Calculations . . . . . . . . . . . . . . . . 5.4 Impact of LOVL and TGe Scaling in TFETs with gate-source overlap 5.4.1 Effect of Overlap Length Scaling on Line Tunneling . . . . 5.4.2 Bulk Ge source Line Tunnel FET (8 nm < TGe ≤ 25 nm) . . 5.4.3 Ultra-Thin Ge Source Line Tunnel FET (TGe ≤ 8 nm) . . . 5.5 Ultra-Thin Ge Source Lateral (P-P) TFET . . . . . . . . . . . . . . 5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 84 85 86 87 88 89 91 92 93 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Modeling of Current and Capacitances including Mobile Channel Charge and Ambipolar Behaviour 96 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 CONTENTS 6.2 6.3 6.4 6.5 7 Device Structure and Simulation Models . . . . . . . . . . . . . . . . Model Description . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3.1 Surface potential . . . . . . . . . . . . . . . . . . . . . . . . 6.3.2 Effect of mobile charges on Φs and k2 : dual modulation effect 6.3.3 Electric Field . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3.4 BTBT Current . . . . . . . . . . . . . . . . . . . . . . . . . 6.3.5 Terminal Charge and Capacitance . . . . . . . . . . . . . . . Model Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.1 Surface Potential, Electric field, and Current . . . . . . . . . . 6.4.2 Terminal Charges and Capacitances . . . . . . . . . . . . . . Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 98 98 102 104 105 107 109 109 112 112 Conclusions 117 7.1 Scope for Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 A List of Publications 120 A.1 Journal Papers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 A.2 Conference Papers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 List of Figures 1.1 (a) The effect of scaling trend on the supply (VDD ) and threshold voltage (Vth ) with CMOS technology generation, reproduced from [8]. (b) The scaling effect on the power density against gate length (LG ). The leakage power density has increased exponentially with LG scaling, reproduced from [9]. . . . . . . . . . . . . . . . . . . . . . 2 1.2 The schematic view of Double Gate Tunnel Field Effect Transistor. . . . . . . . . . . 7 1.3 The schematic cross section view of TFET and MOSFET, reproduced from [46]. The injection of carrier mechanism from the source to channel is thermionic emission in MOSFET, while in TFET’s, it is interband tunneling. The range of energy over which tunneling takes place forms an energy window. The carriers in this energy range tunnel into the empty states of the channel. . . . . . . . . . . . . . . . . . . . . . . . . . 7 The qualitative comparison and analysis between TFETs, MOSFETs and other semiconductor device. The transfer characteristics is shown for the different device structure, which are engineered to impove the electrostatics of the conventional Si MOSFET (red), multigate device (MuG, blue), III-V and SiGe high-mobility channel (purple), and a TFET (green). TFET has a steep OFF to ON transition and lowest OFF state current. We have shown two point A and B on the transfer characteristics, where the trend is opposite. At point A, TFETs offers best subthrehold swing and an improved ION /IOF F as compared to the MOSFET. However at point B, the current is severely degraded as compared to the other device architecture, reproduced from [35]. . . . . . 9 1.4 1.5 (a) DC Characteristics of an Inverter. (b) Transient analysis of an Inverter. . . . . . . 11 1.6 Transient Analyis of NAND gate. . . . . . . . . . . . . . . . . . . . . . . . . . . 12 1.7 Transient Analysis of NOR gate. . . . . . . . . . . . . . . . . . . . . . . . . . . 12 1.8 Transient Analysis of XOR gate. . . . . . . . . . . . . . . . . . . . . . . . . . . 13 1.9 SR Flip-Flop implementation using TFETs . . . . . . . . . . . . . . . . . . . . . 13 1.10 D Flip-Flop implementation using TFETs . . . . . . . . . . . . . . . . . . . . . . 14 2.1 2.2 The schematic cross sectional view of DMG-DGTFET: (a) with hetero dielectric gate material, and (b) with SiGe source and Hf O2 dielectric. . . . . . . . . 26 Method to deposit two contacting gate material using Tilt angle evaporation and Normal evaporation adapted from Ref. [1] . . . . . . . . . . . . . . . . . . . . 27 xii LIST OF FIGURES 2.3 2.4 2.5 2.6 xiii (a) DMG structure reported in Ref. [3] (Laux = 30 nm, Ltun = 20 nm and TSiO2 = 3 nm). (b) Drain to source current in the structure shown in Fig. 2.3a for different sets of BTBT parameters. The BTBT parameters that are considered in Ref. [3] severely overestimates the BTBT current with theoretical/experimental BTBT parameters [14, 15]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 (a) Comparison of SGTFET with DGTFET with SiO2 dielectric (L= 25 nm, tox = 2 nm and φm = 4.2 eV).(b) DGTFET (L=25 nm and tox = 2 nm) based on SiO2 +Hf O2 hetero dielectric gate for different value of metal workfunction (φm ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 Transfer characteristics and energy band diagram for DMG-DGTFET based on hetero dielectric gate material (φtun = 4.0 eV, Ltun = 7nm and Laux = 18 nm). (a) Transfer characteristics at VDS = 1.0 V shown on both semi log and linear scale. Energy band diagram along a horizontal cut line near to the surface is shown in (b) OFF state (VGS = 0.0 V and VDS = 1 V), and (c) ON-state (VGS =1.5 V and VDS = 1 V). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 Transfer characteristics and energy band diagram for DMG-DGTFET based on hetero dielectric gate material (φaux = 4.4 eV, Ltun = 7 nm and Laux = 18 nm). (a) Transfer characteristics at VDS = 1.0 V shown on both semi log and linear scale. Energy band diagram along a horizontal cut line near to the surface is shown in (b) OFF state (VGS = 0.0 V and VDS = 1 V), and (c) ON-state (VGS =1.5 V and VDS = 1 V). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.7 (a) Transfer characteristics for mono and hetero dielectric DMG-DGTFET (φaux = 4.4 eV and φtun =4.0 eV) at VDS = 1.0 V. (b) Energy band diagram along a horizontal cut line near to the surface shown at (VGS = -0.5 V and VDS = 1.0 V). . . . 34 2.8 (a)Transfer characteristics with respect to drain voltage for different gate voltages in DMG-DGTFET based on hetero dielectric material gate (φaux = 4.4 eV, φtun =4.0 eV, Ltun = 7 nm and Laux = 18 nm). (b) Band-to-Band generation rate at the surface for DMG-DGTFET based on hetero gate dielectric for VGS = 1.0, 0.9, 0.8 and 0.7 V, respectively at VDS =1.0 V. . . . . . . . . . . . . . . . . . . 35 DMG-DGTFET based on SiGe source at 50 % Ge mole fraction in SiGe, and hafnium oxide gate dielectric (Ltun = 5 nm and Laux = 20 nm). Transfer characteristics at VDS = 1.0 V shown on both semi log and linear scale, (a) for φtun = 4.0 eV and different value of φaux , and (b) for φaux = 4.4 eV and different value of φtun . (c) Output characteristics at φaux =4.4 eV and φtun =4.0 eV for different values of gate voltage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 2.10 DMG-DGTFET based on SiGe source at 50 % Ge mole fraction in SiGe, and different gate dielectric material (Ltun = 5 nm and Laux = 20 nm). (a) Energy band diagram along a horizontal cut line near to the surface at ON-state (VGS =1.5 V VDS = 1V). (b) Transfer characteristics at φaux = 4.4 eV and φtun = 4.0 eV for different dielectrics. . . . . . . . . . . . . . . . . . . . . . . . . . . 37 2.9 LIST OF FIGURES 2.11 The effect of different Ge mole fraction on DMG-DGTFET performance with hafnium oxide gate dielectric and hetero SiGe on the source. (a) Energy band diagram is shown at ON state (VGS = 1.5 V and VDS = 1.0 V). (b) Transfer characteristics on log and linear scale at φtun =4.0 eV and φaux =4.4 eV, and (c) plot of average subthreshold slope against Ge mole fraction in SiGe. (d) DIBL effect evaluated at different Ge mole fraction with φtun =4.0 eV and φaux = 4.4 eV. . . 3.1 3.2 3.3 3.4 3.5 3.6 3.7 xiv 38 Schematic of DGTFET with compressively strained SiGe (c-SiGe) source and hetero dielectric gate: (a) without pocket implant, and (b) with pocket implant. (c) Extended SiGe/Si/SiGe source structure with mono dielectric gate. . . . . . 46 Qualitative depiction of the strain effects in Si and SiGe following Ref. [3]. (a) The effect of compressive strain in SiGe showing the hydrostatic strain shift and uniaxial component of strain. (b) Valence band splitting of V1 and V2 band for different mole fraction of Ge in c-SiGe and t-Si. . . . . . . . . . . . . . . . . . 49 DGTFET (L = 25 nm and tox =2 nm) based on the c-SiGe source, 0.5 Ge mole fraction, and HD gate for different values of φm . (a) Transfer characteristics at VDS = 1.0 V shown on both semi log and linear scale. Energy band diagram along a horizontal cut line near to the surface is shown at:(b) ON-state (VGS =1.5 V and VDS = 1V), and (c) OFF-state (VGS = 0.0 V and VDS = 1V). (d) Output characterisitcs for different gate voltages. . . . . . . . . . . . . . . . . . . . . 54 DGTFET based on c-SiGe source, 0.5 Ge mole fraction and different hetero gate dielectric. (a) Energy band diagram calculated along a horizontal plane near to the surface is shown at ON state. (b) Transfer characteristics at φm =4.2 eV at VDS =1.0 V. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 The effect of different mole fraction of Ge(x) in DGTFET based on c-SiGe with HD gate. (a) Energy band diagram along a horizontal cut line near to the surface is shown at ON-state (VGS =1.5 V and VDS = 1 V). (b) Transfer characteristic at VDS = 1.0 V at φm =4.2 eV. (c) SSAV G and ION v/s Ge mole fraction (x). . . . . 56 DGTFET (L = 25 nm and tox =2 nm) based on the c-SiGe source, 0.5 Ge mole fraction, HD gate and pocket implant near the source junction. (a) Energy band diagram along a horizontal cut line near the surface is shown at ONstate (VGS =1.5 V and VDS = 1 V) for different pocket implant doping (NP OC ) at φm =4.2 eV. (b) Transfer characteristics at VDS = 1.0 V is shown on both semi log and linear scale for different value of NP OC . (c) SSAV G and ION against NP OC doping. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 Transfer characteristics of DGTFET with extended c-SiGe source and 0.5 Ge mole fractions for different encroachment length (nm) inside the gate. As the source is extended inside the gate, tunneling is delayed due to the increase in tunnel barrier width because of the depletion of the extended part of the source. (b) Subthreshold slope and ON-current for different extension length inside the gate. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 LIST OF FIGURES 3.8 3.9 Energy band diagram along a horizontal cut near to the surface calculated at: (a) ON state (VGS =1.5 V and VDS =1 V). The tunnel barrier width increases with the extension length and 2nm extension gives the best BTBT rate.(b) OFF state (VGS =0 V and VDS =1 V). No tunneling path is formed for any of the structures in the OFF state as shown by dotted line. . . . . . . . . . . . . . . . . . . . . . xv 59 (a) Transfer characteristics of DGTFET with extended c-SiGe source, relaxed SiGe source (r-SiGe) (Ge mole fraction = 0.5) and all Si TFET. (b) The effect of FIQC in tunneling showing the current components and the shift in the tunneling. 60 3.10 Channel length scaling of DGTFET (a) with gate aligned structure, and (b) Source extended structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 4.1 Schematic of the double gate TFET with extended Ge source and hetero-dielectricgate (HG). We consider a device of channel length 40 nm, n-type drain doping of 1018 cm−3 , p-type source doping of 1020 cm−3 , and body is low n-type doped (1016 cm−3 ), HG gate dielectric with 2 nm thickness. Source is extended inside the gate by 2 nm. Work function of metal at the gate is Φm = 4.2 eV . . . . . . . 67 4.2 The confined bandstructure for: (a) Ge 4 nm film and (b) Si 6 nm thin film. The confinement increases the band gap in both cases. Note that Si becomes direct band gap semiconductor at reduced thickness [23]. . . . . . . . . . . . . . . . 69 (a) Thickness scaling of total BTBT current considering direct and indirect valley of Si (VDS = 0.7 V). Subsurface BTBT improves the current in subthreshold region with thickness scaling of Ge films due to the improved gate coupling. (b) Variation of ON current with Ge thickness for a constant gate overdrive voltage of 0.8 V. Ge film thickness scaling results in decrease of ION (due to increase in band gap) and improvement in average subthreshold slope (because of the improved gate coupling). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 Energy band diagram along a horizontal cut line at distance of 1Å from the surface (at VDS = 0.7 V) for: (a) different values of gate voltage at 4 nm thin film and (b) at OFF state (VGS = 0V) for different ultra thin thickness. In (a) as gate voltage increases tunneling shifts from Ge-Si junction to Ge. In (b) as thickness is scaled, the barrier gets wider due to the confinement effect on Si and Ge. The voltage range over which tunneling takes place from Ge to Si increases with thickness scaling, because of the change in band alignment. . . . 75 Transfer characteristics of DGTFET with extended Ge source, where the total current is broken down into its constituent components of BTBT and SRH current at 4 nm thickness. At OFF state (VGS = 0V), the total current consists of both SRH and BTBT current. . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 4.3 4.4 4.5 LIST OF FIGURES 4.6 5.1 5.2 5.3 5.4 xvi BTBT generation rate contour for different VGS at VDS = 0.7 V for different thickness. At VGS = 0.2 V, the BTBT generation rate is higher for 6 nm thickness, which is due to the simultaneous effect of the improved electrostatics of the system and better BTBT parameter value. At low VGS , there is tunneling from Ge to Si. For VGS = 0.6 V, the BTBT generation rate of 8 nm thin film is higher as compared to the other thickness. This is due to the improved gate coupling with VGS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 Schematic of a double gate TFET with gate-source overlap, n-type drain doping of 1 × 1018 cm−3 , p-type source doping of 1 × 1020 cm−3 , and low n-type body doping of 1 × 1016 cm−3 . The gate dielectric is HfO2 with 2 nm thickness. The nominal gate-source overlap length (LOVL ) and the channel length are 8 nm and 40 nm respectively. The metal gate work function is set to 4.2 eV. . . . . . . . 86 Transfer characteristics of DGTFET with different gate-source overlap lengths (LOVL ). The vertical line tunneling increases with LOVL , because of increased band bending provided by the overlapped part. Inset shows the ION and SSAVG variation with LOVL . ION increases linearly with LOVL . SSAVG remains almost constant with LOVL , but sharply increases at very low LOVL , because of increase in OFF current. SSAVG is calculated as the inverse of slope between VT (VGS at which IDS = 1 × 10−7 A/µm) and VOFF . . . . . . . . . . . . . . . . . . . . . 88 (a) Transfer characteristics of DGTFET with gate-source overlap, LOVL = 8 nm. Lateral P-P tunneling takes place initially at low VGS , and line tunneling sets in at higher VGS . Both, lateral and vertical tunneling will contribute to the total current. The inset shows ION and SSAVG for thickness scaling from bulk to ultra-thin Ge source. The ON current and SSAVG essentially remain constant as thickness is scaled down from bulk to 8 nm. For Ge thickness less than 8 nm, the ON current and SSAVG degrade, because of the disappearance of line tunneling (see Fig. 5.3a for more details).(b) Energy band diagrams are shown for different values of VGS at VDS = 0.7 V for 8 nm thickness along the lateral direction considered at 1 Å from the interface. Bands are pulled down with increase in VGS , which improves lateral P-P tunneling.(c) Energy band diagram along the vertical direction. As VGS increases, bands are pulled down and at around 0.5 V, vertical tunneling path is formed in the device, and this initiates line tunneling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 Transfer characteristics for ultra-thin body Ge source having thickness less then 8 nm. The current degrades in the line tunneling configuration as thickness is scaled to 4 nm. Inset shows the energy band diagram along the vertical direction for ultra-thin Ge source at VGS = 1 V and VDS = 0.7 V. Gate voltage is not sufficient to provide enough band bending needed to initiate line tunneling in the device, and hence the total current is reduced. . . . . . . . . . . . . . . . . 90 LIST OF FIGURES 5.5 6.1 6.2 6.3 6.4 6.5 6.6 Transfer characteristics of DGTFET with lateral P-P Tunneling. There is a cross over point in the transfer characteristics, which is due to the effect of gate coupling and gate screening. Inset shows the structure of lateral P-P TFETs. Source is extended inside the gate by 2 nm to limit the tunneling in Ge only. The hetero dielectric gate is composed of SiO2 (LSiO2 = 35 nm) and HfO2 (LHfO2 = 5 nm). ION and SSAVG variation with thickness is shown in the inset. ION degrades at reduced thickness of Ge source because of gate screening. . . . . . . . . . . . xvii 91 Schematic of the DGTFET used in the study, where R1 is the source depletion region, R2 is the channel region, and R3 is the drain depletion region. VG , VS , and VD are the applied gate, source, and drain voltage, respectively and Φi represents the potential at the various marked positions. The doping in various regions and the channel length are mentioned in the figure. The workfunction of the metal at the gate is Φm = 4.2 eV. . . . . . . . . . . . . . . . . . . . . . . 97 Plot of the channel mid potential as a function of VGS for different values of VDS . The linear portion corresponds to the depletion in the channel, while the saturated region indicates inversion in the channel. The model results are consistent with the TCAD simulation data. . . . . . . . . . . . . . . . . . . . . . . . . . 102 Energy band profile in DGTFET showing minimum and maximum tunneling path at; a) source /body junction and b) drain/ body junction. . . . . . . . . . . 106 Validation of the surface potential and electric field behavior of model against the TCAD simulation data along the lateral distance in x-axis. In panel(a), the effect of mobile charge carrier is not included, while it is included in panels (b), (c) and (d). In panel (a), model without mobile charge effect shows excellent matching with simulation data at lower VGS but start deviating with increase in mobile charge carrier at higher gate voltage at VGS =1.2V. The model in depletion condition at VGS =0.2V in panel (b) has excellent match with simulation data. The deviation in panel (a) at higher VGS highlights the necessity to include the mobile charge carriers in the model, which corrects the deviation at high VGS values as shown in (c). The panel (c) displays the surface potential profile along the channel for VGS =1.0V and VDS =0.0V to 1.2V in steps of 0.2V. The saturation in the mid potential is effectively captured by the model along with the pinning of the surface potential at the source end. (d) Validation of the total electric field predicted by the model along the channel with TCAD simulation data in the channel for VDS = 1.0V and different VGS values with smoothing function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 Drain current predicted by the model verified with TCAD simulation data; (a) transfer characteristics at varying gate voltage for different VDS . (b) Output current characteristics at different VGS values, perfect saturation in the drain current is achieved due to the pinning of the surface potential at the source side. 111 Terminal charge (QD , and QG ) predicted by the model against simulation data; (a) Drain terminal charge for different VDS w.r.t. applied gate to source voltage and (b) Gate terminal charge for different VDS w.r.t. applied gate to source voltage.111 LIST OF FIGURES 6.7 xviii Validation of drain capacitance (CDG ) and gate capacitance (CGG ) behavior of model against the TCAD simulation data. The panel (a) shows the CDG while panel (b) shows the CGG at different values of VDS w.r.t. applied gate to source voltage. Note that CGG and CDG looks similar, but their values are different as QG includes QS in addition to the QD . . . . . . . . . . . . . . . . . . . . . . . 112 Chapter 1 Introduction The revolution in integrated circuit (IC) technology has led to a new era in the field of electronics with tremendous computational capabilities and computing power. The research and development carried out by the electronics community over the past four decades has resulted in the continued miniaturization of the conventional MOSFET dimensions leading to larger circuit density and higher functionality at reduced cost. Gordon Moore (Intel co-founder) predicted that “over the history of computing hardware, semiconductor chips would become two times more powerful every two years at lower cost, and they would ultimately be so small, that they could be embedded in homes, cars and smartphones” [1]. The predictions of Moore proved to be accurate for several decades, and guided/motivated the semiconductor industry to use them as benchmarks to set the goals for research and development. The evolution in IC technology and digital electronics is strongly linked to Moore’s law: quality adjusted microprocessor prices, memory capacity sensors and even the number and size of pixels in digital cameras. The miniaturization of the conventional MOSFET according to Moore’s law, cannot be continued forever. In 2015, Gordon Moore perceived that the pace of progress would eventually reach saturation, which started appearing at 22 nm node in 2012, and is continuing at 14 nm [1, 2]. Later Intel made an announcement that “our cadence today is closer to two and a half years than two”. This is scheduled to hold through the 10 nm width in late 2017 [2, 3]. They also made an announcement that as we progress from 14 nanometer technology node to 10 nanometer and plan for 7 nanometer and 5 nanometer and even beyond, our plans are proof that Moores law is still applicable [4–6]. In order to continue with Moore’s law, today’s hardware 1 1.1 CMOS Crisis 2 (a) (b) Figure 1.1: (a) The effect of scaling trend on the supply (VDD ) and threshold voltage (Vth ) with CMOS technology generation, reproduced from [8]. (b) The scaling effect on the power density against gate length (LG ). The leakage power density has increased exponentially with LG scaling, reproduced from [9]. needs to utilize many architecture cores, and the programming should be done in a threaded manner to take the full advantage. 1.1 CMOS Crisis The scaling of conventional MOSFET has started reaching saturation as already discussed. The continuous miniaturization of the complementary metal-oxide semiconductor (CMOS) fieldeffect transistors (FETs) has resulted in extraordinary improvements in the switching speed, transistor density, functionality and reduced cost of microprocessors. But on the flip side, miniaturization has resulted in increased power density on the chip, which starts affecting the reliability of chips. Importance of this observation can be realized from the fact that todays CMOS microprocessor, operates around the power density of a nuclear reactor [7]. The power crisis in the conventional MOSFET arises due to the fundamental reason that the supply voltage, applied to operate the transistors has not scaled proportionately with transistor density. Alternatively it can be said that although the transistors dimensions have continued to shrink, but the supply voltage used to trigger these transistors doesn’t vary with technology generation as 1.1 CMOS Crisis 3 shown in Fig. 1.1a. The supply voltage is nearly flattened at approximately 1 V for 90 nm node and beyond [8]. The saturation in supply voltage scaling is attributed to the non-scalability of the conventional MOSFET threshold voltage. The supply voltage scaling reduces the dynamic power consumption, which is desirable for power constrained applications. But to achieve the same ON-current, threshold voltage Vth should scale proportionately, as VDD reduces. The reduction in Vth causes an exponential increase in OFF-current and hence the static leakage power. The exponential increase in OFF-current is due to subthreshold swing limit of 60 mV/decade, which in turn is due to the Boltzmann distribution of carriers at room temperature. Hence the transistor dissipates power, even when it is supposed to be off (see Fig. 1.1b) [9]. In order to know the speed of transition from OFF to ON state, subthreshold slope is defined. The subthreshold slope is used as a performance parameter for the device and it can be defined as point to point subthreshold slope and average subthreshold slope [10]: Point subthreshold slope is measured as the amount of voltage that is required to change the current by one order of magnitude, when the transistor is operated in the subthreshold regime and mathematically expressed as: S= ∂ log ID ∂ log VG −1 ∂VG ∂ϕS ∂ϕS ∂ log ID Cdep kT kT = 1+ ln10 ≥ ln10 . Cox q q = (1.1) In Eq. 1.1, the first term on right hand side is the gate coupling efficiency, and it is a measure of gate coupling to channel potential, and is greater than 1 for conventional MOSFET. The second term is due to the thermal Boltzmann distribution of mobile carriers in the channel and is fundamentally limited to 60 mV/dec at room temperature. Thus from Eq. 1.1, it can be inferred that even if the gate is coupled perfectly to the surface potential so that 1 + Cdep /Cox = 1, or Cox >> Cdep , it requires minimum of 60 mV gate voltage to bring one order of change in the current magnitude. This sets the fundamental limit on the subtheshold slope and shows why Vth cannot be scaled continuously in a conventional MOSFET. The off current, IOF F , increases exponentially with reduction in Vth , hence leakage power dissipation also increases exponentially [11]. On the other hand, if Vth is set high, then the excessive gate overdrive 1.2 Alternative Solutions to Scale CMOS 4 voltage would not be sufficient to drive the device at high speed [11]. Thus a trade off between power (energy) dissipation and speed (delay), sets a fundamental energy efficiency limit for CMOS technology. Another way to define the steepness is through the average subthreshold slope, SSAV G , and it is calculated as the slope in the ID − VG between Vth (threshold voltage) and Vof f (gate off voltage) as: SSAV G = Vth − Vof f , log IVth − log IVof f (1.2) where IVth and IVof f are the drain current at Vth and Vof f . The electrostatic control of gate on the channel degrades as the channel shrinks, due to the increase in closeness between source and drain junction [12]. Also the gate coupling to various sub-surfaces reduces with scaling, resulting in the increased off state leakage current [13]. Additionally, the drain control on channel increases with scaling and hence the leakage current due to the drain voltage also increases, and this is typically referred as drain induced barrier lowering (DIBL) [14]. In order to deal with these new challenges to scaling, various solutions like retrograde substrate doping profile, shallow source/drain extensions, halo implants and strained silicon [15, 16] have been proposed. Adoption of these solutions for the scaling of planar MOSFETs makes the fabrication process more complex [17]. A promising alternative to scale CMOS further, is to choose new MOSFET architectures eg. Fully depleted SOI (FD-SOI) [18] and FinFET [19, 20] and new channel materials eg. MoS2 [21], III-V [22] and Graphene [23]. 1.2 Alternative Solutions to Scale CMOS In order to circumvent the challenges associated with the conventional MOSFET scaling, new ideas and phenomena are being explored to come up with the new possibilities. The increasing difficulties in Si CMOS scaling has created the need of investigation of alternative channel materials, novel device architectures, and new phenomenon are being investigated and these are presented in the following sections. 1.2 Alternative Solutions to Scale CMOS 1.2.1 5 Alternative Channel Materials New materials are being explored to replace Si in the conventional MOSFET. Materials like germanium (Ge) [24] and III-V alloy semiconductors [25] have opened a new portal to a new era of ultra-low-power and high-speed applications. Further, layered 2-D materials have initiated another area of immense possibilities. Materials like Graphene, TMD materials (M oS2 etc), and Phosphorene etc have shown excellent material properties and have supported electric fieldcontrolled bandgap tuning [26]. The problem with the 2-D material is that their interface with the dielectric is not good. High quality gate dielectric is required to sustain a very high vertical electric fields in such devices. The fabrication of devices using these materials is a challenge, as it can be either n type (III-V compound semiconductor) device or p type (Ge) device but not both, which are required for CMOS circuit designs. The quantum confinement effects, which are dependent on material thickness scaling can be used as a powerful techniques to optimize the material properties, because materials having different thicknesses can be viewed as different materials with entirely different properties. This sets a new platform for electronics and optoelectronic applications. 1.2.2 Alternative Device Architectures The new device architectures like fully depleted SOI (FD-SOI, partially depleted PD-SOI) and FinFET are being used in sub-micron technolody nodes for silicon technology . The device is structured in such a manner, so that the electrostatics of system is improved. These new device architectures provide better gate control on the channel, as channel is surrounded by the gate from multiple sides [27]. Thickness can be scaled in these devices without increasing subsurface leakage paths and, hence, unwanted leakage currents are suppressed [28]. The mobility is also improved because of undoped body, which eliminates surface roughness scattering [11]. To adjust the threshold voltage, channel doping is not used as a parameter in these structures. Thus, random dopant fluctuations, a major cause of process variability in bulk MOSFETs, are eliminated [29–31]. FinFET is another novel device architecture to deal with MOSFET scaling issues. It has 1.3 Tunnel Field Effect Transistor 6 a thin-body non-planar MOSFET constructed on silicon-on-insulator (SOI) or bulk substrate, using patterning and etching technologies. In FinFET, the SS is improved (i.e., close to 60 mV/decade at room temperature) and SCE are suppressed by optimizing the fin size [14]. The massive scaling is possible for FinFET and FD-SOI, while maintaining good device performance [32]. These technologies are now being widely used in Integrated Circuits (ICs) due to lower power consumption. Gate all around FETs (GAAFETs) are also being widely explored. They provide excellent electrostatic control. However, the stacked-Nanowire NW-GAA FET technology involves fabrication complexities, like formation of high aspect ratio fin structure etc [33]. Recently, a new multi-gate transistor architecture having oxide inserted FinFET (iFinFET), has been presented to overcome the fabrication issues. The iFinFET has better gate control over the channel as the fringing electric field enhances the coupling between the gate and channel regions [34]. 1.2.3 Alternative Device Phenomena The devices, which are based on some new transport phenomenan are being explored to replace conventional MOSFET. Tunnel field effect transistor (TFET) [35] and impact ionization MOSFET (IMOS) [36, 37] are some of the devices, which use new phenomena band to band tunneling and impact ionization respectively, for carrier transport. TFETs have been widely researched recently for low power VLSI devices, because of their exceptional leakage power performance and low OFF current. IMOS suffers from the high voltage requirement required for the impact ionization [37]. The limitations with TFETs are their low ON-current (ION ) compared to ITRS requirements and stringent fabrications steps required for extracting performance [35]. In our work, we have carried out thorough investigation of Tunnel Field Effect Transistor, as mentioned in the following sections. 1.3 Tunnel Field Effect Transistor Tunnel field effect transistor (TFET), because of its low subthreshold swing and excellent reduction in leakage current [38–41] is attracting attention of the electronics community. They 1.3 Tunnel Field Effect Transistor 7 Figure 1.2: The schematic view of Double Gate Tunnel Field Effect Transistor. Figure 1.3: The schematic cross section view of TFET and MOSFET, reproduced from [46]. The injection of carrier mechanism from the source to channel is thermionic emission in MOSFET, while in TFET’s, it is interband tunneling. The range of energy over which tunneling takes place forms an energy window. The carriers in this energy range tunnel into the empty states of the channel. are being investigated for various low power applications [42–44]. TFET is a gated control PIN diode, in which tunneling takes place from valence band of source to the conduction band of the channel and this is controlled by the gate voltage (see Fig. 1.2). The gate voltage is used to modulate the tunneling barrier width at the source/body junction. The tunneling phenomenon was first observed in Tunnel diode by Esaki [45] in 1960. TFET can be operated at much lower voltage than the conventional MOSFET and, hence it consumes less power than the conventional MOSFET [42], but operating mechanisms in TFET and MOSFET are quite different. 1.3.1 MOSFET vs TFET The structural difference between the conventional MOSFET and TFET is the asymmetric source and drain junctions as shown in Fig. 1.3. TFET can be viewed as a gated tunnel diode 1.3 Tunnel Field Effect Transistor 8 in series with the conventional MOSFET connected via an internal node. The carrier injection mechanism from source to channel is different in both the devices as shown in Fig. 1.3. Unlike the conventional MOSFET, where the injection of carrier takes place over the potential barrier (i.e.thermionic emission), the carriers in TFETs are injected into the channel through the potential barrier via a process called band-to-band tunneling (BTBT) (see Fig. 1.3) [47, 48]. The transfer characteristics of TFET are also quite different from conventional MOSFET. TFET has some offset with respect to the drain voltage in the transfer characteristics [49, 50]. The operation of conventional MOSFET relies on the modulation of channel potential by the applied gate voltage for the injection of carriers from the source into the channel through a process called thermionic emission. The carriers are generated thermally and are distributed in the source region with the Boltzmann distribution, and the resulting current can be expressed as: IT hermionic VGS ∼ . = exp nVT (1.3) The barrier at the source/channel junction is modulated by the applied gate voltage. In OFF state, the barrier seen by the carrier from the source region is large, hence resulting current is leakage current (see Fig. 1.3). The current is not zero because the thermal (exponential) distribution of carriers still permits a finite number of carriers to be injected over the large potential barrier. In the ON state, as the gate voltage increases the barrier at the source/channel junction is modulated and this results in exponential injection of carriers from the source into the channel (exponential current modulation). The thermal (Boltzmann) distribution of carriers in the source region ultimately defines, how steeply a MOSFET can switch from on to off-state. The steepness, defined by the amount of voltage required to modulate the current by one-order of magnitude, is called the subthreshold swing (S) and is limited to 60 mV/dec at room temperature [10]. In Fig. 1.4, we have shown the qualitative comparison of the transfer characteristics for TFETs and other device architectures for MOSFETs [35]. TFET shows better subthrshold swing as compared to the MOSFETs. However, the ON current associated with the TFETs is 1.4 Operation of Tunnel Field Effect Transistor 9 Figure 1.4: The qualitative comparison and analysis between TFETs, MOSFETs and other semiconductor device. The transfer characteristics is shown for the different device structure, which are engineered to impove the electrostatics of the conventional Si MOSFET (red), multigate device (MuG, blue), III-V and SiGe high-mobility channel (purple), and a TFET (green). TFET has a steep OFF to ON transition and lowest OFF state current. We have shown two point A and B on the transfer characteristics, where the trend is opposite. At point A, TFETs offers best subthrehold swing and an improved ION /IOF F as compared to the MOSFET. However at point B, the current is severely degraded as compared to the other device architecture, reproduced from [35]. very less as compared to the ITRS requirements. 1.4 Operation of Tunnel Field Effect Transistor The operation of TFET relies on the gate modulation of channel potential and the alignment of band at the source/channel junction. The range of energy over, which bands are aligned at the source/channel junction, forms an energy window (band pass filter) over which tunneling takes place as shown in Fig. 1.3. The path connecting two points of the same energy level is referred as a tunnel path. It may be noted here that the minimum tunneling path dominates the BTBT current. The alignment of bands at the junction can be staggered band type-I or type-II or broken band gap alignment. The BTBT current in a TFET can be toggled ON and OFF by controlling the band bending through the applied gate voltage. TFET is an ambipolar device that can show p-type behavior 1.4 Operation of Tunnel Field Effect Transistor 10 with dominant hole conduction and n-type behavior with dominant electron conduction. By using an asymmetric doping profile at the source and drain junction or by using a hetero junction, the tunneling at the drain end can be suppressed and the ambipolar current can be reduced. The asymmetry in TFET also achieves a low off-state current. In OFF state, the valence band and the conduction band are not aligned, as a result no tunneling path is formed in the device. Hence BTBT is suppressed and the OFF state current is limited by the reverse leakage current. As the gate voltage increases in the negative direction, bands are pulled upward. The conduction band gets aligned with the valence band and a conducting channel is formed from the source valence band to the channel conduction band. At low temperature, the energy distribution of carriers at the source is restricted and the higher energy part of the source fermi level is effectively cut off. This filtering function makes it possible to achieve an subthreshold slope of below 60mV per decade [35, 51]. The key to better voltage scaling of a TFET than a MOSFET is that subthreshold slope remains below 60mV per decade over several orders of magnitude of drain current. 1.4.1 Shortcomings of the TFETs The main stumbling blocks with TFET are its low ON state current (ION ) compared to ITRS requirements and complicated fabrication steps required for extracting good performance. The low ON current in TFETs results in increased delay associated with the circuits. The methods to increase the ON current in TFETs can be categorised as follows: Structural modifications (hetero junction source [52, 53] etc): The low ON current issue in TFETs can be tackled by structural modifications. By using a hetero junction, with a smaller band gap material at the source, the ON current can be boosted. Raised Ge source TFETs [54], gate source overlapped TFETs (Vertical tunneling TFETs [55]), extended source TFETs [52], and L-shapted TFETs [56] have been used to increase the ON current. In a conventional lateral TFET, tunneling takes place in the lateral direction from the source to the channel. This tunneling is referred to as point to point-(P-P) lateral tunneling. However, if sufficient gate to source 1.4 Operation of Tunnel Field Effect Transistor (a) 11 (b) Figure 1.5: (a) DC Characteristics of an Inverter. (b) Transient analysis of an Inverter. overlap is provided, then the BTBT can be initiated in the vertical direction as well from the bulk to surface. This tunneling is referred to as line tunneling or vertical tunneling [55]. Doping optimization (retrograde doping, pocket implant source [57, 58] etc): The BTBT takes place within few nm along side the source/body junction. This BTBT is a very strong function of the steepness in the energy bands, which can be controlled by the doping profile in the device. Typically n+ pocket implant near the source is utilized to improve the ON current in TFETs. However, the OFF current also increases by a large amount due to the SRH electric field enhanced recombination. Gate metal work function engineering (double gate, dual metal gate [59, 60]): The workfunction of the metal at gate can be engineered to optimize the performance of TFETs. The steepness of energy bands at source/body junction can be controlled by the metal workfunction. The techniques like Dual metal gate can be utilized to individually control the different portion of the transfer characteristics. Different types of contacts can also be used at the source (eg. wrap around contacts) to play with the metal workfunction. Improving electrostatics (gate all around structures, nano wires [55]): Improving the electrostatics is another option to increase the ON current in TFETs. In Gate all around (GAA) TFET and nano wires, the gate electrostatic control on the tunnel junction is improved, which increases the BTBT current. Dielectric optimization (with one or multiple dielectrics etc. [61–63]): High-k dielectric material at the gate results in better coupling of the gate to the tunnel junction. Mono and hetero 1.5 Circuits Performance Using InAs HomoTFETs (a) (b) 12 (c) Figure 1.6: Transient Analyis of NAND gate. (a) (b) (c) Figure 1.7: Transient Analysis of NOR gate. dielectric gate materials can be employed at the gate to boost the ON current in TFETs. In hetero dielectric gate, the OFF current can be reduced further as comprared to the mono dielectric. Strain Engineering is another way to boost the performance of TFETs. The strain at source and drain ends can be used to increase the ON current [64–66]. 1.5 Circuits Performance Using InAs HomoTFETs To assess the circuits performance of TFETs with the best state of art ON current, we have simulated circuits using 20nm InAs homoTFETs. The III-V Tunnel FET Model is a lookup table based model, where the device current and capacitance characteristics are obtained from experimentally calibrated TCAD Sentaurus simulation [67]. The detailed experimental calibration and the look up based model of TFET is described in Ref. [67]. The results thus obtained for logic gates are shown in Fig. 1.5 for an inverter, Fig. 1.6 for 1.5 Circuits Performance Using InAs HomoTFETs (a) 13 (b) (c) Figure 1.8: Transient Analysis of XOR gate. (a) (b) (d) (c) (e) Figure 1.9: SR Flip-Flop implementation using TFETs NAND gate, Fig. 1.7 for NOR Gate, and Fig. 1.8 for XOR gate, respectively. The sequential circuits are also simulated using look-up-based model. The results for SR and D Flips are shown in Fig. 1.9 and Fig. 1.10, respectively. From these figures, it is clear that TFET based circuits can operate at low VDD . 1.6 Thesis Goals and Outline (a) 14 (b) (c) (d) Figure 1.10: D Flip-Flop implementation using TFETs 1.6 Thesis Goals and Outline The goal of this thesis is to get insight into the underlying physics of the TFETs. The main stumbling blocks with TFET are its low ON current and complicated fabrication steps, which restrict its usefulness for main stream device. The various optimization techniques are adopted to boost the performance of TFET as already briefly discussed above. In the rest of the chapters, we will discuss some of the optimization technique in details. The first principle density functional theory (DFT) thickness dependent quantum confinement effects on Ge source extended lateral TFET is also discussed for lateral and vertical tunneling. We have also presented a surface potential based analytical model for Si-DGTFET accounting the mobile charges in the channel. The thesis is organized as follows: In Chapter 2, we have discussed “Dual Metal Gate” (DMG) technique in detail for the mono and hetero dielectric gate with Si and relaxed SiGe source. The hetero dielectric at the gate results in better subthreshold slope as compared to the mono dielectric. DMG with SiGe at 1.6 Thesis Goals and Outline 15 the source allows the tunable band gap and results in improved performance as compared to the Si TFETs. The effect of varying the mole fraction of Ge in SiGe and the short channel effects are also discussed for this structure. In Chapter 3, strain engineering is discussed for the SiGe source. We have incorporated the effect of the compressive strain at the source in SiGe and tensile strain in Si for DGTFET with strained SiGe source. Sandwitched SiGe/Si/SiGe source is presented, which results in the compressive strain at the source. The lattice constant of pseudomorphically grown SiGe on Si substrate is higher as a result compressive strain is developed at the source. Further, at the vertical interface between SiGe/Si source, Si channel is tensile strained. The effect of the pocket implant at the source is also discussed. Additionally source extended structure is discussed to find the optimized source extension length. In Chapter 4, the extended Ge source lateral tunnel field effect transistor is scrutinized. The direct and indirect valley in Germanium (Ge) are separated by a very small offset, which opens up the prospect of direct tunneling in the Γ valley of extended Ge source TFET. We explore the impact of thickness scaling of extended Ge source lateral TFET on the band to band tunneling (BTBT) current. Ge source is extended inside the gate by 2 nm to confine the tunneling in Ge only. We observe that as thickness is scaled, the band alignment at Si/Ge hetero junction changes significantly, which results in increase in Ge to Si BTBT current. Based on density functional calculations we first obtain the bandstructure parameters (bandgap, effective masses etc.) for the Ge and Si slabs of varying thickness, and these are then used to obtain the thickness dependent Kane’s BTBT tunneling parameters. The Kane’s BTBT parameters of Si and Ge are then used to obtain the scaled BTBT tunneling current. The electrostatics improves as thickness is reduced in the ultra thin Ge flim (≤ 10 nm). The ON current degrades as we scale down in thickness, however the subthreshold slope (SSAVG ) improves remarkably with thickness scaling due to subsurface BTBT. In Chapter 5, we have presented a comparative study of vertical and lateral tunneling in Ge source TFETs as the source is scaled down from bulk to ultra-thin slabs of 4 nm thickness. We have theoretically calculated BTBT parameters for ultra-thin thickness using band structure results from atomistic simulations, while for bulk we have considered bulk parameters. In the 1.6 Thesis Goals and Outline 16 investigated double gate TFET (DGTFET) structures, we have used an overlapped gate over Ge source to include vertical line tunneling along with lateral tunneling in the device. First, the impact of the gate overlap length scaling has been investigated at constant Ge source thickness of 8 nm. We have then analyzed the impact of scaling the Ge source thickness keeping the gate overlap length fixed. We have observed that Ge source thickness scaling till 8 nm has negligible effect on the ON current, but as the thickness is further scaled down to ultra-thin Ge films, line tunneling disappears in the device. We also report TFETs with ultra-thin Ge source, but small gate-source overlap to confine the tunneling in Ge only. In this configuration, lateral tunneling drives the drain current, leading to better sub-threshold slope and ON current. In Chapter 6, we have developed a surface potential based analytical model for double gate tunnel field effect transistor (DGTFET) for the current, terminal charges, and terminal capacitances. The model accounts for the effect of the mobile charge in the channel and captures the device physics in depletion as well as in the strong inversion regime. The narrowing of the tunnel barrier in the presence of mobile charges in the channel is incorporated via modeling of the inverse decay length, which is constant under channel depletion condition and bias dependent under inversion condition. To capture the ambipolar current behavior in the model, tunneling at the drain junction is also included. The proposed model is validated against TCAD simulation data and it shows close match with the simulation data. Finally in, Chapter 7 summarizes the research work carried out in this thesis and also suggests the scope of future work in this area. Bibliography [1] G. Moore, “Gordon Moore: The Man Whose Name Means Progress. The visionary engineer reflects on 50 years of Moore’s Law,” IEEE Spectrum Special Report, Mar 2015. [2] “INTEL CORP, FORM 10-K (Annual Report), Filed 02/12/16 for the Period Ending 12/26/15.” [Online]. Available: https://files.shareholder.com/downloads/INTC/ 867590276x0xS50863-16-105/50863/filing.pdf [3] D. Clark, “Intel Rechisels the Tablet on Moore’s Law,” Wall Street Journal Digits Tech News and Analysis, Jul 2015. [Online]. Available: https://blogs.wsj.com/digits/2015/07/ 16/intel-rechisels-the-tablet-on-moores-law/ [4] T. Bradshaw, “Intel chief raises doubts over Moore’s law,” Financial Times, Jul 2015. [Online]. Available: https://www.ft.com/content/36b722bc-2b49-11e5-8613-e7aedbb7bdb7 [5] R. Waters, “As Intel co-founder’s law slows, a rethinking of the chip is needed,” Financial Times, Jul 2015. [Online]. Available: https://www.ft.com/content/ 4d8dabaa-2bd5-11e5-acfb-cbd2e1c81cca [6] J. Niccolai, “Intel pushes 10nm chip-making process to 2017, slowing Moore’s Law,” Info World, Jul 2015. [Online]. Available: https://www.pcworld.idg.com.au/article/579795/ intel-pushes-10nm-chip-making-process-2017-slowing-moore-law/ [7] “Intel Processors–Specifications.” [Online]. Available: http://ark.intel.com/Default.aspx [8] P. Packan, “Device and Circuit Interactions,” IEEE International Electron Device Meeting Short Course: Performance Boosters for Advanced CMOS Devices, Dec 2007. 17 BIBLIOGRAPHY 18 [9] B. Meyerson, (IBM)Semico Impact Conference, Taiwan, Jan 2004. [10] S. M. Sze and K. K. Ng, “Physics Of Semiconductor Devices.” Wiley: Hoboken, New Jersey, 2007, pp. 315–316. [11] C. C. Hu, “Modern Semiconductor Devices for Integrated Circuits.” Prentice Hall: Upper Saddle River, New Jersey, 2010, pp. 281–283. [12] P. K. Chatterjee and J. E. Leiss, “An analytic charge-sharing predictor model for submicron MOSFETs,” in IEEE International Electron Devices Meeting, 1980, pp. 28–33. [13] D. Neamen, “Semiconductor Physics And Devices.” McGraw-Hill, Inc.: New York, 2003. [14] J. P. Colinge, “FinFETs and Other Multi-Gate Transistors (Integrated Circuits and Systems).” Springer Science and Business Media: US, 2008. [15] S. Thompson, “MOS Scaling: Transistor Challenges For the 21st Century,” Intel Technology Journal, 1998. [16] Y. Zhang, H. Zhu, H. Wu, Y. Zhang, Z. Zhao, J. Zhong, H. Yang, Q. Liang, D. Wang, J. Li, C. Jia, J. Liu, Y. Zhao, C. Li, L. Meng, P. Hong, J. Li, Q. Xu, J. Gao, X. He, Y. Lu, Y. Zhang, T. Yang, Y. Wang, H. Cui, C. Zhao, H. Yin, H. Zhong, H. Yin, J. Yan, W. Wang, D. Chen, H. Yu, S. Yang, and T. Ye, “Planar bulk MOSFETs with self-aligned pocket well to improve short-channel effects and enhance device performance,” IEEE Transactions on Electron Devices, vol. 62, no. 5, pp. 1411–1418, 2015. [17] P. Magarshack, P. Flatresse, and G. Cesana, “UTBB FD-SOI: A process/design symbiosis for breakthrough energy-efficiency,” in Proc. of Design, Automation Test in Europe Conference Exhibition, pp. 952–957, 2013. [18] Q. Xie, C. Jung Lee, J. Xu, C. Wann, J. Sun, and Y. Taur, “Comprehensive Analysis of Short-Channel Effects in Ultrathin SOI MOSFETs,” IEEE Transactions on Electron Devices, vol. 60, no. 6, pp. 1814–1819, 2013. BIBLIOGRAPHY 19 [19] E. Amat, A. Calomarde, C. Almudever, N. Aymerich, R. Canal, and A. Rubio, “Impact of FinFET and IIIV/Ge Technology on Logic and Memory Cell Behavior ,” IEEE Transactions on Device and Material Reliability, vol. 14, no. 1, pp. 344–350, 2014. [20] N. Paydavosi, S. Venugopalan, Y. S. Chauhan, J. P. Duarte, S. Jandhyala, A. M. Niknejad, and C. Hu, “BSIM-SPICE models enable FinFET and UTB IC designs,” IEEE Access, vol. 1, no. 3, pp. 201–215, 2013. [21] K. Majumdar, C. Hobbs, and P. D. Kirsch, “Benchmarking Transition Metal Dichalcogenide MOSFET in the Ultimate Physical Scaling Limit,” IEEE Electron Device Letters, vol. 35, no. 3, pp. 402–404, 2014. [22] N. Agrawal, Y. Kimura, R. Arghavani, and S. Datta, “Impact of Transistor Architecture (Bulk Planar, Trigate on Bulk, Ultrathin-Body Planar SOI) and Material (Silicon or IIIV Semiconductor) on Variation for Logic and SRAM Applications ,” IEEE Transactions on Electron Devices, vol. 60, no. 10, pp. 3298–3304, 2013. [23] F. Schwierz, “Graphene-based FETs,” in Proc. of Ninth International Conference on Advanced Semiconductor Devices and Microsystems, pp. 131–138, 2012. [24] R. Pillarisetty, “Academic and industry research progress in germanium nanodevices,” Nature, vol. 479 7373, pp. 324–328, 2011. [25] J. Del Alamo, “Nanometre-scale electronics with III-V compound semiconductors,” Nature, vol. 479 7373, pp. 317–323, 2011. [26] T. Chu and Z. Chen, “Bandgap engineering in 2D layered materials,” in Proc. of Technical Digest of IEDM, pp. 707–710, 2015. [27] J. G. Fossum and V. P. Trivedi, “Fundamentals of Ultra-Thin-Body MOSFETs and FinFETs.” Cambridge University Press, 2013. BIBLIOGRAPHY 20 [28] Y. K. Choi, K. Asano, N. Lindert, V. Subramanian, T. J. King, J. Bokor, and C. Hu, “Ultrathin-body SOI MOSFET for deep-sub-tenth micron era,” IEEE Electron Device Letters, vol. 21, no. 5, pp. 254–256, 2000. [29] M. H. Chiang, J. N. Lin, K. Kim, and C. T. Chuang, “Random dopant fluctuation in limited-width FinFET technologies,” IEEE Transactions on Electron Devices, vol. 54, pp. 2055–2060, 2007. [30] S. Markov, A. Zain, B. Cheng, and A. Asenov, “Statistical variability in scaled generations of n-channel UTB-FD-SOI MOSFETs under the influence of RDF, LER, OTF and MGG,” in Proc. of SOI Conference (SOI), pp. 1–2, 2012. [31] Y. Yang, S. Markov, B. Cheng, A. Zain, X. Liu, and A. Cheng, “Back-gate bias dependence of the statistical variability of FDSOI MOSFETs with thin BOX,” IEEE Transactions on Electron Devices, vol. 60, no. 2, pp. 739–745, 2013. [32] “FinFET technology market 2016.” [Online]. Available: http://www.researchandmarkets. com/research/pflc9w/finfet technology [33] T. J. K. Liu, “Sustaining the silicon revolution from 3-D transistors to 3-D integration,” SPIE Advanced Lithography 2015 Conference, Mar. 2015. [34] P. Zheng, D. Connelly, F. Ding, and T. J. K. Liu, “Considerations for ultimate CMOS scaling,” IEEE Transactions on Electron Devices, vol. 62, no. 12, pp. 3945–3950, 2015. [35] A. M. Ionescu and H. Riel, “Tunnel field-effect transistors as energy-efficient electronic switches,” Nature, vol. 479, no. 7373, pp. 329–337, Nov 2011. [36] K. Gopalakrishnan, P. B. Griffin, and J. D. Plummer, “Impact Ionization MOS (I-MOS)Part I : Device and Circuit Simulations,” IEEE Transactions on Electron Devices, vol. 52, no. 1, Jan 2005. BIBLIOGRAPHY 21 [37] K. Gopalakrishnan, R. Woo, C. Jungemann, P. B. Griffin, and J. D. Plummer, “Impact ionization MOS (I-MOS)-Part II : Experimental Results,” IEEE Transactions on Electron Devices, vol. 52, no. 1, Jan 2005. [38] Q. Zhang, W. Zhao, and A. Seabaugh, “Low-subthreshold-swing tunnel transistors,” IEEE Electron Device Letters, vol. 27, no. 4, pp. 297–300, Apr 2006. [39] K. Gopalakrishnan, P. B. Griffin, and J. D. Plummer, “I-MOS: a novel semiconductor device with a subthreshold slope lower than kT/q,” in IEEE International Electron Devices Meeting, Dec 2002, pp. 289–292. [40] M. Luisier and G. Klimeck, “Performance comparison of tunneling field-effect transistors made of InSb, Carbon, and GaSb-InAs broken gap heterostructures,” in IEEE International Electron Devices Meeting, Dec 2009, pp. 1–4. [41] J. Singh, K. Ramakrishnan, S. Mookerjea, S. Datta, N. Vijaykrishnan, and D. Pradhan, “A novel Si-Tunnel FET based SRAM design for ultra low-power 0.3 V VDD applications,” in 15th Asia and South Pacific Design Automation Conference (ASP-DAC), Jan 2010, pp. 181–186. [42] U. E. Avci, R. Rios, K. J. Kuhn, and I. A. Young, “Comparison of power and performance for the TFET and MOSFET and considerations for P-TFET,” in 11th IEEE International Conference on Nanotechnology, Aug 2011, pp. 869–872. [43] J. Nez and M. J. Avedillo, “Comparison of TFET and CMOS Using Optimal Design Points for Power Speed Tradeoffs,” IEEE Transactions on Nanotechnology, vol. 16, no. 1, pp. 83–89, Jan 2017. [44] N. Gupta, A. Makosiej, A. Vladimirescu, A. Amara, and C. Anghel, “Ultra Low Power Compact TFET Flip Flop Design for High Performance Low Voltage Applications,” in 17th International Symposium on Quality Electronic Design (ISQED), Mar 2016, pp. 107– 112. BIBLIOGRAPHY 22 [45] R. Noyce, “A 50- year old notebook reveals the seed of a great invention,” IEEE Spectrum Special Report, May 2005. [46] S. Datta, “Interband Tunnel Transistor: Oppurtunities and Challenges.” University Lec- ture: Electrical Engineering, University of Notre Dame, Notre Dame, 2015. [47] E. Kane, “Zener tunneling in semiconductors,” Journal of Physics and Chemistry of Solids, vol. 12, no. 2, pp. 181 – 188, 1960. [48] E. O. Kane, “Theory of Tunneling,” Journal of Applied Physics, vol. 32, no. 1, pp. 83–91, 1961. [49] C. Shen, S. L. Ong, C. H. Heng, G. Samudra, and Y. C. Yeo, “A Variational Approach to the Two-Dimensional Nonlinear Poisson’s Equation for the Modeling of Tunneling Transistors,” IEEE Electron Device Letters, vol. 29, no. 11, pp. 1252–1255, Nov 2008. [50] K. Boucart and A. M. Ionescu, “A new definition of threshold voltage in Tunnel FETs,” Solid-State Electronics, vol. 52, no. 9, pp. 1318 – 1323, 2008. [51] J. Knoch, S. Mantl, and J. Appenzeller, “Impact of the dimensionality on the performance of tunneling FETs: Bulk versus one-dimensional devices,” Solid-State Electronics, vol. 51, no. 4, pp. 572 – 578, 2007. [52] M. Zhu, S. S. Tan, E. H. Toh, and E. Quek, “Method of Fabricating a Silicon Tunneling Field Effect Transistor (TFET) with High Drive Current,” Apr 14 2011, US Patent 12 587 511. [53] P. Y. Wang and B. Y. Tsui, “Band Engineering to Improve Average Subthreshold Swing by Suppressing Low Electric Field Band-to-Band Tunneling With Epitaxial Tunnel Layer Tunnel FET Structure,” IEEE Transactions on Nanotechnology, vol. 15, no. 1, pp. 74–79, Jan 2016. BIBLIOGRAPHY 23 [54] S. H. Kim, S. Agarwal, Z. A. Jacobson, P. Matheu, C. Hu, and T. J. K. Liu, “Tunnel Field Effect Transistor With Raised Germanium Source,” IEEE Electron Device Letters, vol. 31, no. 10, pp. 1107–1109, Oct 2010. [55] J. Schulze, A. Blech, A. Datta, I. A. Fischer, D. Hhnel, S. Naasz, E. Rolseth, and E.M. Tropper, “Vertical Ge and GeSn heterojunction gate-all-around tunneling field effect transistors,” Solid-State Electronics, vol. 110, no. Supplement C, pp. 59 – 64, 2015. [56] Z. Yang, “Tunnel Field-Effect Transistor With an L-Shaped Gate,” IEEE Electron Device Letters, vol. 37, no. 7, pp. 839–842, Jul 2016. [57] D. B. Abdi and M. J. Kumar, “In-Built N Pocket p-n-p-n Tunnel Field-Effect Transistor,” IEEE Electron Device Letters, vol. 35, no. 12, pp. 1170–1172, Dec 2014. [58] P. Jain and D. Kumar, “Drive Current Boosting Using Pocket Implant Near to the Strained SiGe/Si Source with Single-Metal/Dual-Metal Double-Gate Tunnel Field-Effect Transistor,” in Proceeding of International Conference on Intelligent Communication, Control and Devices : ICICCD 2017. Springer Singapore, 2017, pp. 943–950. [59] J. Knoch, “Optimizing Tunnel FET performance - Impact of Device Structure, Transistor Dimensions and Choice of Material,” in International Symposium on VLSI Technology, Systems, and Applications (VLSI-TSA), 2009, pp. 45–46. [60] K. H. Kao, A. S. Verhulst, W. G. Vandenberghe, B. Soree, W. Magnus, D. Leonelli, G. Groeseneken, and K. D. Meyer, “Optimization of Gate-on-Source-Only Tunnel FETs With Counter-Doped Pockets,” IEEE Transactions Electron Devices, vol. 59, no. 8, pp. 2070–2077, Aug 2012. [61] M. J. Lee and W. Y. Choi, “Effects of Device Geometry on Hetero-Gate-Dielectric Tunneling Field-Effect Transistors,” IEEE Electron Device Letters, vol. 33, no. 10, pp. 1459– 1461, Oct 2012. BIBLIOGRAPHY 24 [62] R. Gandhi, Z. Chen et al., “Vertical Si-Nanowire n-Type Tunneling FETs With Low Subthreshold Swing (≤ 50 mV/decade) at Room Temperature,” IEEE Electron Device Letters, vol. 32, no. 4, pp. 437–439, Apr 2011. [63] P. Jain, V. Prabhat, and B. Ghosh, “Dual metal-double gate tunnel field effect transistor with mono/hetero dielectric gate material,” Journal of Computational Electronics, vol. 14, no. 2, pp. 537–542, Jun 2015. [64] T. Krishnamohan, D. Kim, S. Raghunathan, and K. Saraswat, “Double-Gate StrainedGe Heterostructure Tunneling FET (TFET) With Record High Drive Currents and < 60mV/dec Subthreshold Slope,” in Proceedings IEEE International Electron Devices Meeting, 2008, pp. 1–3. [65] O. M. Nayfeh, C. N. Chleirigh, J. Hennessy, L. Gomez, J. L. Hoyt, and D. A. Antoniadis, “Design of Tunneling Field-Effect Transistors Using Strained-Silicon/StrainedGermanium Type-II Staggered Heterojunctions,” IEEE Electron Device Letters, vol. 29, no. 9, pp. 1074–1077, Sept 2008. [66] K.-H. Kao, A. S. Verhulst, R. Rooyackers, B. Douhard, J. Delmotte, H. Bender, O. Richard, W. Vandervorst, E. Simoen, A. Hikavyy, R. Loo, K. Arstila, N. Collaert, A. Thean, M. M. Heyns, and K. D. Meyer, “Compressively strained SiGe band-to-band tunneling model calibration based on p-i-n diodes and prospect of strained SiGe tunneling field-effect transistors,” Journal of Applied Physics., vol. 116, pp. 214 506–1–214 506–11, 2014. [67] H. Liu, V. Saripalli, V. Narayanan, and S. Datta, “III-V Tunnel FET Model 1.0.1.” NanoHub: Penn State University, 2015. Chapter 2 Dual Metal - Double Gate Tunnel Field Effect Transistor 2.1 Introduction In the previous chapter, we have discussed the need for a new device, which can circumvent the issues of the conventional MOSFET. We have proposed that TFET can offer a promising solution at reduced operating voltage. In this chapter, to tackle the issues of conventional Si TFETs, Dual Metal-Double Gate Tunnel Field Effect Transistor (DMG-DGTFET) is discussed for mono and hetero dielectric gate material (HD). The DMG technique offers the flexibility to have low as well as high work function metals at the gate to control the individual portion of the transfer characteristics. This results in pragmatic increase in ON current, improved subthreshold slope and reduced threshold voltage. Additionally, the DMG technique is used along with the mono/hetero dielectric gate material to optimize the performance of Si based TFET. The hetero dielectric that we have used at the gate is a combination of SiO2 and Hf O2 . The 2-D device simulations are performed using Sentaurus TCAD and the results thus obtained are discussed using energy band diagram, tunneling barrier width. The comparison has also been made between mono and hetero dielectric gate. On application of hetero dielectric to the Si DMGDGTFET, the advantages of both the techniques combines, and it results in higher ION /IOF F (2 × 109 ) compared to the mono dielectric case (2.5 × 108 ). The average subthreshold slope also improves from 58 mV/decade in mono dielectric to 48 mV/decade in hetero dielectric Si 25 2.1 Introduction 26 (a) (b) Figure 2.1: The schematic cross sectional view of DMG-DGTFET: (a) with hetero dielectric gate material, and (b) with SiGe source and Hf O2 dielectric. DMG-DGTFET. Further, to address the low ON current issue in Si TFETs, DMG-DGTFET with type-II staggered band alignment SiGe hetero junction source is presented. The good compatibility of SiGe with Si leads to better quality hetero junction SiGe source. The hetero SiGe source offers tunable band gap with variation in Ge mole fraction and results in improved device transfer characteristics. All the simulations are done in Synopsys TCAD for a channel length of 25nm using the non-local BTBT tunneling model. The previous reported work on DMG-DGTFET in Ref. [1–3] have discussed the, Dual Metal Gate technique for Si, in detail. However, there is no discussion on ambipolarity, gate leakage, and impact of different dielectric materials [2, 3]. The rigorous discussion of different dielectrics on DGTFET is presented in Ref. [4]. The concept of hetero dielectric is explored in details in Ref. [5, 6]. To the best of our survey, nobody has reported DMG-DGTFET with hetero dielectric gate and relaxed SiGe source. DMG-DGTFET with vertical tunneling has also not been reported. The rest of the chapter is organized as follows: In section-2.2, we have elaborated the structure of Si DMG-DGTFET [2, 7] with mono and hetero-dielectric gate [5, 8] and DMG-DGTFET with SiGe source. The fabrication details and the simulation model that are incorporated are also discussed in this section. In section-2.3, double gate Si structure is discussed for the hetero dielectric material system. The detailed study of Si based DMG-DGTFETS with hetero /mono dielectric gate is carried out in section 2.4. This is followed by the discussion of DMG-DGTFET with SiGe source in section 2.5. In section 2.6, conclusions are drawn. 2.2 DMG-DGTFET: Schematic, fabrication and simulation. 27 Figure 2.2: Method to deposit two contacting gate material using Tilt angle evaporation and Normal evaporation adapted from Ref. [1] 2.2 DMG-DGTFET: Schematic, fabrication and simulation. The Si DMG-DGTFET based on hetero dielectric material consists of gate with two different metals as shown in Fig. 2.1a. The metal gate which is near to the drain is referred as auxiliary gate, and it can be used to control the off state current by tunning its work function. The metal gate near to the source side is referred as tunnel gate, and it controls the tunneling at the body/source junction. Hetero dielectric material consists of a combination of different dielectric. The dielectric below the tunnel gate has a high dielectric constant, which assists in improving the ON current. The schematic of DMG-DGTFET with SiGe source is shown in Fig. 2.1b, which has SiGe as source and high-k dielectric as gate material. The fabrication process of TFETs using SiGe source and drain has already been suggested in Ref. [9]. The process is similar to the conventional CMOS with some additional steps to deposit the two different work-function metals. Tilt angle evaporation [1, 10] can be used to deposit gate material-1 with a photoresist layer as the shadow and gate material-2 can be deposited using normal evaporation techniques as shown in Fig. 2.2 adapted from Ref. [1]. To form n+ drain region, ion implantation of arsenic (As) is done into Si followed by annealing. The source is formed by etching of Si underneath the gate followed by regrowth of in-situ doped p+ Si1−x Gex [11–13]. The HD gate can be formed in two steps [6]. First, SiO2 gate dielectric is selectively etched away at the source side by using buffered hydrogen fluoride (BHF) solution, 2.2 DMG-DGTFET: Schematic, fabrication and simulation. 28 (a) (b) Figure 2.3: (a) DMG structure reported in Ref. [3] (Laux = 30 nm, Ltun = 20 nm and TSiO2 = 3 nm). (b) Drain to source current in the structure shown in Fig. 2.3a for different sets of BTBT parameters. The BTBT parameters that are considered in Ref. [3] severely overestimates the BTBT current with theoretical/experimental BTBT parameters [14, 15]. while SiO2 gate dielectric at the drain end is protected by photo resist masks. In the next step Hf O2 is deposited by atomic layer deposition of high-k material to fill the gap. The inner high-k spacer is also formed during this step. Then, the outer SiO2 spacer is formed in order to implement gate underlap structures for performance improvement [5, 8]. All the simulations are done using Synopsys TCAD, using the non-local tunneling model [16] for the calculation of BTBT tunneling current. The values of the non-local BTBT tunneling model for Si are calibrated to the experimental results [14] and for SiGe default BTBT parameter values are considered. Phonon assisted tunneling is used in the model. The gate tunneling current is simulated using Direct tunneling model and the traps are considered as fixed charge with concentration of 1010 cm−2 . The SRH recombination model along with the Fermi-Dirac statistics is used throughout the simulation. The doping dependence, high field saturation, and 2.3 DGTFET based on hetero dielectric gate material 29 enormal mobility models are used with their default parameter values [17, 18]. The quantum effects due to the charge centroid shift is considered using the MLDA model in the simulation. Traps are likely to exist at the interface of the gate and body, and are modelled by considering them to be acceptor types, loacted at 0.3 eV from the middle of the band gap [19]. For multiple defects the couple defect level (CDL) recombination is used. A very fine mesh is used at the surface to simulate the structure. The doping profile is assumed to be constant in the drain, source and body region [20, 21]. It is to be mentioned over here that in Sentaurus, the quantum models [3] are not coupled with the non local BTBT tunneling model. Including the confinement effects in BTBT will lead to a VT shift due to the increase in band gap, because of quantization of energy states at the surface. To calibrate our results, we have done the Sentaurus simulation for the structure reported in Ref. [3]. The results thus obtained are shown in Fig. 2.3. We find that the values of the BTBT parameters considered in Ref. [3] severely overestimate the BTBT current for Si as reported in Ref. [14, 15]. Thus, in our work we have considered theoretical/experimental reported values of BTBT parameters from Ref. [14, 15]. The following parameter value have been used throughout this work, drain doping 1018 cm−3 (Arsenic), source doping 1020 cm−3 (Boron), and body is lightly doped with arsenic concentration of 1016 cm−3 . The thickness of the gate dielectric material is 2 nm. Hetero dielectric that we have used in this study is composed of silicon dioxide (SiO2 ) and hafnium oxide (Hf O2 ). The gate length that we used throughout this work is 25 nm. The length of auxiliary gate is Laux = 18 nm and that of tunnel gate is Ltun = 7 nm. The silicon body thickness tSi is 8 nm. The analysis of DMG-DGTFETS has been done for VDD =1.0 V. 2.3 DGTFET based on hetero dielectric gate material The results obtained from simulations of SGTFET and DGTFET with SiO2 dielectric are shown in Fig. 2.4a. As expected DGTFET results in better performance as compared to SGTFET due to better electrostatics of the system. The DGTFET based on hetero dielectric material is simulated and the results are shown in Fig. 2.4b. The analysis is carried out for DGTFET based on SiO2 2.4 Workfunction engineered DGTFET with HD gate (a) 30 (b) Figure 2.4: (a) Comparison of SGTFET with DGTFET with SiO2 dielectric (L= 25 nm, tox = 2 nm and φm = 4.2 eV).(b) DGTFET (L=25 nm and tox = 2 nm) based on SiO2 +Hf O2 hetero dielectric gate for different value of metal workfunction (φm ). and Hf O2 hetero dielectric to find the optimum value of the work function of the metal (φm ). It can be observed from Fig. 2.4b, that the best possible value of ION /IOFF is achieved at φm = 4.3 eV. The purpose of using hetero dielectric is to optimize the OFF-current as compared to the case of a mono dielectric device/TFET. The high-k dielectric is manoeuvre at the source side, since it improves the electrostatics around the tunnel junction. The dielectric on the drain side affects only the OFF-current. Important thing to note here is that by changing the φm in the DGTFET based on hetero dielectric gate simply shifts the transfer characteristics parallel to the X-axis. As a result sub-threhold slope does not change. 2.4 Workfunction engineered DGTFET with HD gate In this section, DMG-DGTFET based on HD gate material is analyzed. The workfunction of the metal at the auxiliary and tunnel gate are engineered following the approach of Ref. [22]. The dual metal at the gate in DMG-DGTFET, along with different dielectrics provide additional flexibility to control different section of the transfer characteristics, thus improving the TFET performance. 2.4 Workfunction engineered DGTFET with HD gate (a) 31 (b) (c) Figure 2.5: Transfer characteristics and energy band diagram for DMG-DGTFET based on hetero dielectric gate material (φtun = 4.0 eV, Ltun = 7nm and Laux = 18 nm). (a) Transfer characteristics at VDS = 1.0 V shown on both semi log and linear scale. Energy band diagram along a horizontal cut line near to the surface is shown in (b) OFF state (VGS = 0.0 V and VDS = 1 V), and (c) ON-state (VGS =1.5 V and VDS = 1 V). 2.4.1 Workfunction engineered - Auxiliary Gate The workfunction at the auxiliary gate (φaux ) is engineered at constant φtun =4.0 eV. At OFF state, as φaux increases, the OFF state current first decreases and then increases as shown in Fig. 2.5a. The OFF state current is composed of reverse leakage current plus the BTBT current. As φaux increases up to 4.4 eV reverse leakage current will dominate, since there is no tunneling path formed in the device. But with further increase in φaux , tunneling path is formed at the drain end as shown in Fig. 2.5b, which increases the OFF state current. At ON state, increase in φaux does not bring any significant change in the tunnel barrier width as shown in Fig. 2.5c. 2.4 Workfunction engineered DGTFET with HD gate (a) 32 (b) (c) Figure 2.6: Transfer characteristics and energy band diagram for DMG-DGTFET based on hetero dielectric gate material (φaux = 4.4 eV, Ltun = 7 nm and Laux = 18 nm). (a) Transfer characteristics at VDS = 1.0 V shown on both semi log and linear scale. Energy band diagram along a horizontal cut line near to the surface is shown in (b) OFF state (VGS = 0.0 V and VDS = 1 V), and (c) ON-state (VGS =1.5 V and VDS = 1 V). The ambipolar current increases for negative increase in VG values as shown in Fig. 2.5a. As φaux increases for constant negative VG bands are pulled upwards. As a result tunneling path is formed at the drain end and barrier width reduces, which increases the current. The gate current under both the auxiliary and tunnel gate decreases with increase in φaux for constant φtun = 4.0 eV. For φaux = 4.4 eV and φtun = 4.0 eV, the maximum value of gate current under both auxiliary and tunnel gate is limited to 10−13 A/µm. 2.4 Workfunction engineered DGTFET with HD gate 33 Table 2.1: Comparison of various parameters for DMG (φaux =4.4 eV and φtun =4.0 eV) technique for mono and hetero dielectric. Mono dielectric (Hf O2 ) and Hetero dielectric (SiO2 + Hf O2 ). Technique mono dielectric DMG-DGTFET hetero dielectric DMG-DGTFET 2.4.2 ION (µA/µm) 24.8 24.8 IOF F (A/µm) 9.8 ×10−14 1.2 ×10−14 ION /IOF F 2.5 ×108 2 ×109 SSAV G (mV/decade) 58 48 Workfunction engineered - Tunnel Gate The workfunction at the tunnel gate (φtun ) is engineered at constant φaux =4.4 eV, to ensure low OFF current. As φtun increases the current starts to decrease, as shown in Fig. 2.6a. This is because of increased tunnel barrier width at the source/body junction as displayed in Fig. 2.6b. At OFF state tunnel gate work function does not bring any significant change in tunnel barrier width as shown in Fig. 2.6c. The ambipolar conduction is not affected by φtun , this is because φtun does not bring significant changes in the energy band at the drain end. Thus current remains unaffected by φtun for negative VG . The gate current under both the gate decreases with increase in φtun for constant φaux = 4.4 eV. The maximum gate current under both the gate is limited to 10−13 A/µm for constant φtun = 4.0 eV and φaux = 4.4 eV. The best possible value for the auxiliary gate work function φaux is 4.4 eV and for tunnel gate φtun it is 4.0 eV. The metals which can be used at gates to obtain the desired φaux of 4.4 eV are W, Ta, and Mo etc, and φtun = 4.0 eV are Mo, Ni-Ti, and Sc etc [23, 24]. The average subthreshold slope, which can be used as a performance parameter in comparison of semiconductor devices is defined as, SSAV G = Vt − Vof f . log(IVt ) − log(IVof f ) (2.1) Here, Vt is the threshold voltage, Vof f is the gate voltage from which the drain current starts to take off [22], IVt is the drain current at Vgs = Vt , and IVof f is the drain current of the device at Vgs = Vof f . The SSAV G shows a significant improvement from 58 mV/decade in mono dielectric gate to 48 mV/decade in the hetero dielectric DMG-DGTFET. The results obtained for DMG-DGTFET based on SiO2 + Hf O2 hetero dielectric and Hf O2 2.4 Workfunction engineered DGTFET with HD gate (a) 34 (b) Figure 2.7: (a) Transfer characteristics for mono and hetero dielectric DMG-DGTFET (φaux = 4.4 eV and φtun =4.0 eV) at VDS = 1.0 V. (b) Energy band diagram along a horizontal cut line near to the surface shown at (VGS = -0.5 V and VDS = 1.0 V). mono dielectric gate are compared in Fig. 2.7a. The hetero dielectric at the gate gives better performance as compared to mono dielectric in Si DMG-DGTFET. The value achieved for ION /IOF F is 2.5×108 in the mono dielectric, while it is 2×109 for hetero dielectric gate DMGDGTFET. Also the ambipolar conduction in mono dielectric is more as compared to hetero dielectric. This can be explained as follows: for mono dielectric the voltage drop would be less in Hf O2 as compared to silicon, as a result the band are pulled upward in mono-dielectric device as compared to SiO2 +Hf O2 hetero dielectric as shown in Fig. 2.7b. The results are also summarized in Table. 2.1. We have not compared the SMG with DMG because the bias points of both the techniques are not same. 2.4.3 Output characteristics We now focus on the transfer characteristic with respect to drain voltage as shown in Fig. 2.8a. The saturation voltage (Vdsat ) in the output characteristics are evaluated as the value of drain voltage for which the drain current reaches 95 % of its final value. For Si-DMG-DGTFET based on hetero dielectric gate material the value of Vdsat comes out to be 0.85, 0.81, 0.76, and 0.70 for the following gate voltages: 1, 0.9, 0.8, and 0.7 respectively. For mono dielectric, the value 2.5 DMG-DGTFET with SiGe Source 35 (a) (b) Figure 2.8: (a)Transfer characteristics with respect to drain voltage for different gate voltages in DMGDGTFET based on hetero dielectric material gate (φaux = 4.4 eV, φtun =4.0 eV, Ltun = 7 nm and Laux = 18 nm). (b) Band-to-Band generation rate at the surface for DMG-DGTFET based on hetero gate dielectric for VGS = 1.0, 0.9, 0.8 and 0.7 V, respectively at VDS =1.0 V. of Vdsat is similar to the hetero dielectric case as the value of ION is similar for both the cases. The band-to-band generation rate is simulated using Kane dispersion relationship and WKB approximation [25] and is shown in Fig. 2.8b for different value of gate voltages. With Kane equation, the peak electron generation rate (Gtun.max ), by BTBT is evaluated to be, 3 2 Gtun.max = A Eg 1 e 2 q 2 wtmin −qwtmin 3 BEg2 , (2.2) where wtmin is the minimum tunneling width. As the value of VG increases minimum barrier width reduces at the junction, which increases the BTBT generation rate as shown in Fig. 2.8b. 2.5 DMG-DGTFET with SiGe Source In this section, we extend the discussion on DMG-DGTFET to include SiGe source. In particular, we will focus on the impact of Ge mole fraction variation, and different gate materials on the TFET characteristics. The short channel effect are also discussed in this section. For SiGe DMG-DGTFET, we have considered the following device dimensions: Laux =20 nm (length of auxiliary gate), Ltun =5 nm (length of tunnel gate), tSi =10 nm (Silicon body thickness), and tSiGe =10 nm (SiGe layer thickness). The thickness of auxiliary and tunnel 2.5 DMG-DGTFET with SiGe Source 36 (a) (b) (c) Figure 2.9: DMG-DGTFET based on SiGe source at 50 % Ge mole fraction in SiGe, and hafnium oxide gate dielectric (Ltun = 5 nm and Laux = 20 nm). Transfer characteristics at VDS = 1.0 V shown on both semi log and linear scale, (a) for φtun = 4.0 eV and different value of φaux , and (b) for φaux = 4.4 eV and different value of φtun . (c) Output characteristics at φaux =4.4 eV and φtun =4.0 eV for different values of gate voltage. gates, tox =2 nm with mono dielectric gate material. 2.5.1 Workfunction engineered DMG-DGTFET with SiGe source The workfunction at the auxiliary and tunnel gate are engineered in the same manner as it is carried out for Si DMG-DGTFET in section 2.4. The results thus obtained are shown in Fig. 2.9a for auxiliary gate metal workfunction, and in Fig. 2.9b for tunnel gate metal workfunction, respectively. The best possible value for the auxiliary gate work function φaux comes out to be 4.4 eV and for tunnel gate φtun it is 4.0 eV. For φaux =4.4 eV and φtun =4.0 eV, the maximum 2.5 DMG-DGTFET with SiGe Source 37 Table 2.2: The summarized result in a DG-DMTFET with SiGe source for different gate dielectrics. Gate Dielectric SiO2 Si3 N4 Hf O2 ZrO2 Dielectric Constant 3.9 7.5 22 29 (a) ION (µA/µm) 14 82.3 322 377.4 IOF F (A/µm) 1.2 ×10−15 1.34 ×10−14 8.4 ×10−14 1 ×10−13 SSAV G (mV/decade) 75.6 46.7 22.8 19 (b) Figure 2.10: DMG-DGTFET based on SiGe source at 50 % Ge mole fraction in SiGe, and different gate dielectric material (Ltun = 5 nm and Laux = 20 nm). (a) Energy band diagram along a horizontal cut line near to the surface at ON-state (VGS =1.5 V VDS = 1V). (b) Transfer characteristics at φaux = 4.4 eV and φtun = 4.0 eV for different dielectrics. value of gate current under auxiliary gate is limited to 10−13 A/µm, while for tunnel gate it is limited to 10−12 A/µm. The ON current for the proposed device comes out to be 300 µA/µm. Besides ON current, subthreshold slope is very crucial parameter for defining the device performance. The average subthreshold slope at 50 % Ge mole fraction in SiGe DMG-DGTFET comes out to be 22.8 mV/decade. The drain saturation voltage (Vdsat ) is also computed for Fig. 2.9c. The value of Vdsat comes out to be 0.87 V, 0.85 V, 0.83 V, and 0.80 V for gate voltage of 1.0 V, 0.9 V, 0.8 V, and 0.7 V, respectively. 2.5.2 Effect of gate dielectrics The drain current results obtained for different gate dielectric materials along with the energy band diagram at ON state are shown in Fig. 2.10a. The gate oxide materials used are silicon 2.5 DMG-DGTFET with SiGe Source (a) (c) 38 (b) (d) Figure 2.11: The effect of different Ge mole fraction on DMG-DGTFET performance with hafnium oxide gate dielectric and hetero SiGe on the source. (a) Energy band diagram is shown at ON state (VGS = 1.5 V and VDS = 1.0 V). (b) Transfer characteristics on log and linear scale at φtun =4.0 eV and φaux =4.4 eV, and (c) plot of average subthreshold slope against Ge mole fraction in SiGe. (d) DIBL effect evaluated at different Ge mole fraction with φtun =4.0 eV and φaux = 4.4 eV. dioxide (SiO2 ), silicon nitride (Si3 N4 ), hafnium oxide (Hf O2 ), and zirconium oxide (ZrO2 ) with dielectric constant of 3.9, 7.5, 22, and 29, respectively. Higher dielectric constant of the gate dielectric leads to higher ON current, as shown in Fig. 2.10b. This is a consequence of the reduction in barrier width at the junction, which increases the BTBT tunnelling probability. As evident from Table-2.2, that with increasing dielectric constant of gate materials, sub-threshold slope also improves due to better gate control on the channel with a slight increase in the OFF state current. 2.6 Summary 2.5.3 39 Effect of Ge mole fraction in hetero SiGe source The effect of varying the Ge mole fraction in hetero SiGe layer on the transfer characteristics along with the energy band diagram at ON state is shown in Fig. 2.11. As the Ge mole fraction increases energy bandgap decreases at the source as shown in Fig. 2.11a. This results in increased steepness in the energy band near to the source. The increased steepness in the energy band decreases the tunneling barrier width at the tunneling junction, which increases the BTBT current (see Fig. 2.11b). The Ge mole fraction variation changes the energy band gap, which has little impact on intrinsic carrier density and hence on the OFF state current, as shown in Fig. 2.11b. The average subthreshold slope calculated with variation in Ge mole fraction is displayed in Fig. 2.11c. The impact of Ge mole fraction variation on the drain induced barrier lowering (DIBL) is shown in Fig. 2.11d. The DIBL is calculated as the difference in gate voltage for the drain current of 1 nA/µm from the transfer characteristics at drain voltage of 1.0V and 0.1 V. As the Ge mole fraction increases the barrier width at the junction reduces which increases the BTBT tunnelling current. Therefore, increasing the Ge mole fraction in SiGe reduces the DIBL effect. The use of high-k dielectric further reduces the DIBL effects due to better gate coupling with channel. 2.6 Summary The DMG technique, when applied to Si-DGTFET results in individual control for the different section of the transfer characteristics by adjusting the work function of the two metal. This results in the improvement of the ION /IOF F ratio and the average subthreshold slope. We find that the performance can be improved further by use of a different dielectrics under the gate (hetero dielectric). However the use of high-k dielectric results in the degradation of the mobility at the interface due to columbic scattering. This mobility degradation can be improved by using Retro grade kind of doping profile, where the doping is kept low at the surface and high in the body. Thus we can conclude that DMG along with hetero dielectric provides a very powerful technique to boosts the performance of DMG-DGTFET. One limitation of the DMG- 2.6 Summary 40 DGTFET is the increased fabrication complexity particularly for the hetero dielectric devices, as the device dimensions scale down The DMG technique along with the SiGe source, further boost the ON current as compared to Si based DMG-DGTFETs. Increasing the Ge mole fraction in the SiGe source, also leads to improved subthreshold slope. The device with higher dielectric constant materials show better ON current and further improvement in short channel effects. Bibliography [1] W. Long, H. Ou, J. M. Kuo, and K. K. Chin, “Dual-material gate (DMG) field effect transistor,” IEEE Transactions on Electron Devices, vol. 46, no. 5, pp. 865–870, May 1999. [2] A. Chaudhry and M. J. Kumar, “Investigation of the novel attributes of a fully depleted dual-material gate SOI MOSFET,” IEEE Transactions on Electron Devices, vol. 51, no. 9, pp. 1463–1467, Sept 2004. [3] S. Saurabh and M. J. Kumar, “Novel Attributes of a Dual Material Gate Nanoscale Tunnel Field-Effect Transistor,” IEEE Transactions on Electron Devices, vol. 58, no. 2, pp. 404– 410, Feb 2011. [4] K. Boucart and A. M. Ionescu, “Double-Gate Tunnel FET With High-k Gate Dielectric,” IEEE Transactions on Electron Devices, vol. 54, no. 7, pp. 1725–1733, Jul 2007. [5] B. Bhowmick and S. Baishya, “Hetero Double Gate-dielectric Tunnel FET with record high ION /IOFF ratio,” IJCA Proceedings on International Conference on VLSI, Communications and Instrumentation (ICVCI), no. 4, pp. 11–13, 2011. [6] G. Lee, J.-S. Jang, and W. Y. Choi, “Dual-dielectric-constant spacer hetero-gate-dielectric tunneling field-effect transistors,” Semiconductor Science and Technology, vol. 28, no. 5, p. 052001, 2013. 41 BIBLIOGRAPHY 42 [7] M. J. Lee and W. Y. Choi, “Effects of Device Geometry on Hetero-Gate-Dielectric Tunneling Field-Effect Transistors,” IEEE Electron Device Letters, vol. 33, no. 10, pp. 1459– 1461, Oct 2012. [8] K. Y. Na and Y. S. Kim, “Silicon Complementary Metal Oxide Semiconductor Field Effect Transistors with Dual Work Function Gate,” Japanese Journal of Applied Physics, vol. 45, no. 12R, p. 9033, 2006. [9] A. Murthy, R. S. Chau, T. Ghani, and K. R. Mistry, “Semiconductor transistor having a stressed channel,” Sept. 2003, US Patent 66 211 31. [10] A. Yagishita, T. Saito, K. Nakajima, S. Inumiya, Y. Akasaka, Y. Ozawa, K. Hieda, Y. Tsunashima, K. Suguro, T. Arikado, and K. Okumura, “High performance damascene metal gate MOSFETs for 0.1 um regime,” IEEE Transactions on Electron Devices, vol. 47, no. 5, pp. 1028–1034, May 2000. [11] M. Zhu, S. S. Tan, E. H. Toh, and E. Quek, “Semiconductor transistor having a stressed channel,” Apr 14 2011, US Patent 12 587 511. [12] M. Schmidt, L. Knoll, S. Richter, A. Schafer, J. M. Hartmann, Q. T. Zhao, and S. Mantl, “Si/SiGe hetero-structure tunneling field effect transistors with in-situ doped SiGe source,” in 13th International Conference on Ultimate Integration on Silicon (ULIS), Mar 2012, pp. 191–194. [13] Z. Cheng, J. Jung, M. L. Lee, A. J. Pitera, J. L. Hoyt, D. A. Antoniadis, and E. A. Fitzgerald, “Hole mobility enhancement in strained-Si/strained-SiGe heterostructure pMOSFETs fabricated on SiGe-on-insulator (SGOI),” Semiconductor Science and Technology, vol. 19, no. 5, p. L48, 2004. [14] G. Paasch and H. bensee, “A Modified Local Density Approximation Electron Density in Inversion Layers,” physica status solidi (b), vol. 113, no. 1, pp. 165–178, 1982. BIBLIOGRAPHY 43 [15] K. H. Kao, A. S. Verhulst, W. G. Vandenberghe, B. Soree, G. Groeseneken, and K. D. Meyer, “Direct and Indirect Band-to-Band Tunneling in Germanium-Based TFETs,” IEEE Transactions on Electron Devices, vol. 59, no. 2, pp. 292–301, Feb 2012. [16] C. Lombardi, S. Manzini, A. Saporito, and M. Vanzi, “A physically based mobility model for numerical simulation of nonplanar devices,” IEEE Transactions on Computer-Aided Design of Integrated Circuits and Systems, vol. 7, no. 11, pp. 1164–1171, Nov 1988. [17] W. G. Vandenberghe, B. Sore, W. Magnus, G. Groeseneken, and M. V. Fischetti, “Impact of field-induced quantum confinement in tunneling field-effect devices,” Applied Physics Letters, vol. 98, no. 14, p. 143503, 2011. [18] J. Yipeng, W. Kangliang, W. Taihuan, D. Gang, and L. Xiaoyan, “Comparison of bandto-band tunneling models in Si and SiGe junctions,” Journal of Semiconductors, vol. 34, no. 9, p. 092002, 2013. [19] X. Y. Huang, G. F. Jiao, W. Cao, D. Huang, H. Y. Yu, Z. X. Chen, N. Singh, G. Q. Lo, D. L. Kwong, and M. F. Li, “Effect of Interface Traps and Oxide Charge on Drain Current Degradation in Tunneling Field-Effect Transistors,” IEEE Electron Device Letters, vol. 31, no. 8, pp. 779–781, Aug 2010. [20] S. Saurabh and M. J. Kumar, “Estimation and Compensation of Process-Induced Variations in Nanoscale Tunnel Field-Effect Transistors for Improved Reliability,” IEEE Transactions on Device and Materials Reliability, vol. 10, no. 3, pp. 390–395, Sept 2010. [21] M. N. Darwish, J. L. Lentz, M. R. Pinto, P. M. Zeitzoff, T. J. Krutsick, and H. H. Vuong, “An improved electron and hole mobility model for general purpose device simulation,” IEEE Transactions on Electron Devices, vol. 44, no. 9, pp. 1529–1538, Sept 1997. [22] M. Luisier and G. Klimeck, “Performance comparisons of tunneling field-effect transistors made of InSb, Carbon, and GaSb-InAs broken gap heterostructures,” in IEEE International Electron Devices Meeting (IEDM), Dec 2009, pp. 1–4. BIBLIOGRAPHY 44 [23] M. Hasan, H. Park, H. Yang, H. Hwang, H. S. Jung, and J. H. Lee, “Ultralow work function of scandium metal gate with tantalum nitride interface layer for n-channel metal oxide semiconductor application,” Applied Physics Letters, vol. 90, no. 10, p. 103510, 2007. [24] W. Y. Choi and W. Lee, “Hetero-Gate-Dielectric Tunneling Field-Effect Transistors,” IEEE Transactions on Electron Devices, vol. 57, no. 9, pp. 2317–2319, Sept 2010. [25] E. O. Kane, “Theory of Tunneling,” Journal of Applied Physics, vol. 32, no. 1, pp. 83–91, 1961. Chapter 3 Strain Engineering in SiGe source TFETs 3.1 Introduction The previous chapters motivated the need to boost the ON current in Si TFETs, as it is limited by the large bandgap in Si. Further, we have seen that the ON current can be boosted by DMG technique and by using a relaxed SiGe source (low band gap). In this chapter, we study the effect of compressive strain in SiGe source, and the resulting impact on the BTBT current. We have proposed different double gate hetero device structure TFETs incorporating the strain effects at the source and the channel. In the proposed device structure, DGTFET with typeII staggered band alignment, compressively strained SiGe source, and a portion of a tensile strained Si channel with hetero dielectric gate (HD) material is discussed with the experimental BTBT parameter values. The effect of pocket implant underneath the gate at the source/body junction is also discussed for this structure. The use of pocket implant at the source/body junction increases the ON current, but due to SRH electric field enhanced recombination, the OFF state current also increases by a large amount. In the other proposed device structure, DGTFET with type-II staggered band alignment and SiGe/Si/SiGe extended source is discussed with the experimental BTBT parameter values. The different encroachment of SiGe under the gates were explored to find that optimum performance occurs for 2 nm extension. The effect of field induced sub-band discretization at the oxide/semiconductor interface is considered by back envelope calculation method. In Ref. [1, 2], authors have done experimental and theoretical BTBT parameters calcula45 3.1 Introduction 46 (a) (b) (c) Figure 3.1: Schematic of DGTFET with compressively strained SiGe (c-SiGe) source and hetero dielectric gate: (a) without pocket implant, and (b) with pocket implant. (c) Extended SiGe/Si/SiGe source structure with mono dielectric gate. tions. The deviation between experimental and theoretical parameter is accounted by multiplying it with a constant factor. They [1, 2] have not investigated different structures for TFETs, as we have done in our work. Source extended structure is also not explored incorporating the compressive strain effects in SiGe. Additionally, we have incorporated the tensile strain effect in Si on band alignment [3, 4] and the effect of Field Induced Quantum Confinement [5]. Further, we have shown that due to the electric field enhanced SRH recombination, OFF current increases by large amount in pocket implant TFETs. The rest of the chapter is organized as follows: In section-3.2, we have elaborated the structure of c-SiGe source DGTFET without and with pocket implant and HD gate, and also the source extended structure with mono dielectric gate is discussed. The fabrication details and the simulation model that are incorporated are also discussed in this section. In section-3.3, the compressive strain effect on SiGe is discussed. The Kane BTBT parameters for c-SiGe is discussed in section 3.4. This is followed by the discussion of field induced quantum confinement effect in section 3.5. The results that we have obtained from the simulation for different proposed device structure are discuss in section 3.6. In section 3.7, conclusion are drawn. 3.2 Device structure, fabrication, and simulation models. 3.2 47 Device structure, fabrication, and simulation models. The proposed device structure are shown in Fig. 3.1a and Fig. 3.1b for hetero dielectric gate (HD), without and with pocket implant. The sandwiched SiGe/Si/SiGe hetero junction is proposed at the source, that results in a compressive strain at the source and allows to grow higher Ge mole fraction at the surface of SiGe. Due to the vertical lattice mismatch at SiGe/Si hetero junction, portion of the Si channel is tensile strained. In the other proposed device structure, we have extended the compressively strained SiGe source below the gate. The device structure is shown in Fig. 3.1c, in which SiGe/Si/SiGe source structure is extended into the gate side of the channel. SiGe is primarily used to exploit the higher tunneling rate from the SiGe region. This is due to the large built-in electric field at the junction of SiGe with Si, which is caused by the lower band gap of SiGe. 3.2.1 Fabrication The method to fabricate TFETs using SiGe source and drain has already been demonstrated in Ref. [6]. The SiGe source can be formed by etching of Si underneath the gate [6] and regrowth of in-situ doped p+ Si1−x Gex [7, 8]. The hetero dielectric (HD) gate can be fabricated by the steps reported in Ref. [9]. We have assumed the doping profile to be constant in the drain, source and body region [10]. We consider a device of channel length 25 nm, silicon body thickness of 8 nm, n-type drain doping of 1018 cm−3 , p-type source doping of 1020 cm−3 , and body is low n-type doped of 1016 cm−3 . The length of SiO2 gate is 18 nm and that of Hf O2 is 7 nm. The gate dielectric is 2 nm in thickness. Further, the thickness of SiGe layer is 2 nm and it is doped p-type with a doping concentration of 1020 cm−3 . The work function of metal at the gate (φm ) is 4.2 eV. 3.2.2 Simulation models The simulation are done in Sentaurus from Synopsys TCAD, where all relevant physical effects have been included. The band-to-band-tunneling (BTBT) current was simulated by using 3.3 Strain Analysis 48 non local tunneling model, gate leakage mechanism has been included via the direct tunneling model. Due to the very high local electric field assumed in the drift diffusion model, unrealistic impact ionization sets in the device. To avoid this unrealistic impact ionization and to include non-local electric field effect hydrodynamic transport is considered in the simulation. To account for the very high doping in source, we have included band gap narrowing effects and Fermi Dirac statistics. The calibrated parameter values were used for non local BTBT model for compressive strained SiGe following Ref. [1]. The default parameter values of non-local BTBT model for SiGe in Sentaurus are for relaxed SiGe and is calculated for different mole fraction of Ge by using interpolation, which over estimates the experimental values [2]. MLDA model is used to consider the confinement effects due to the charge centroid shift [5]. The compressive strain between SiGe and Si is due to lattice mismatch creates mainly two effects: (1) it reduces the bandgap of relaxed SiGe [2, 11], and (2) it splits the valence and conduction band. The default band gap of SiGe is changed with the experimental values following Ref. [3]. The splitting of the valence and conduction band is considered by including the deformation potential model. The valence band effective mass are considered from the 6band k.p model following Ref. [12]. There are two tunneling path formed in the device due to the splitting of valence band and it is considered by invoking the second tunneling path in the BTBT model with the calibrated BTBT parameter values. The sub band discretization due to the electric field induced quantum confinement is considered by using a correction factor in the onset voltage. In the next section, we analyze the impact of strain on SiGe and Si. 3.3 Strain Analysis The lattice constant of pseudomorphically grown SiGe [001] on [001] Si substrate is higher than Si, as a result biaxial compressive strain develops in SiGe. Due to the vertical lattice mismatch at source/channel junction (SiGe/Si) a portion of the channel is tensile strained. The strain is compressive at the SiGe source and tensile in the Si channel and it’s effect is discussed below. 3.3 Strain Analysis 49 (a) (b) Figure 3.2: Qualitative depiction of the strain effects in Si and SiGe following Ref. [3]. (a) The effect of compressive strain in SiGe showing the hydrostatic strain shift and uniaxial component of strain. (b) Valence band splitting of V1 and V2 band for different mole fraction of Ge in c-SiGe and t-Si. 3.3.1 Compressive Strain in SiGe The compressive strain in SiGe has a hydrostatic component, which shifts the energy bands and a uniaxial component along [001] which causes the degeneracy in the band [3]. It splits the ∆6 ellipsoidal non degenerate conduction band into ∆4 and ∆2 band as shown in Fig. 3.2a. ∆4 bands are in lower energy as compared to ∆2 bands. The strain also splits the valence band into V1 and V2 at the Γ point. The electron effective mass of the c-SiGe ∆4 valley is independent of strain and Ge content [5]. The hole effective mass is considered by using a 6-band k.p band model with a set of luttinger parameter [13] values. The energy separation between V1 and V2 is small at low mole fraction of Ge (x) [3] and it increases with x as shown in Fig. 3.2b. The separation between V1 and V2 bands is only about 80 meV at x=0.5. Thus, we have considered two BTBT path following Ref. [1, 3]. 3.3.2 Tensile strain in Si At the vertical Si/SiGe interface, due to the difference in lattice constant strain is developed in Si channel, which is tensile [4]. This tensile strain in Si channel does exactly opposite of compressive strain i.e it splits the ∆6 band into ∆2 and ∆4 , but the energy of ∆2 band is lower than ∆4 . We have assumed the tensile region to be of 2 nm. Note that increasing the length of this region is not going to affect the tunneling, since tunneling is taking place within few 3.4 BTBT Model Parameters 50 nanometer along the side of the c-SiGe/t-Si junction. For t-Si, we have considered the Si BTBT parameters and band gaps are considered following Ref. [2]. In the following section, we discuss the calibrated values of the BTBT model parameters that are used throughout the simulation. 3.4 BTBT Model Parameters The BTBT current is defined as the integral of the generation rate over the entire device region [14]. Mathematically, Z Z Itun = q GBTBT dydx , (3.1) where GBTBT is the BTBT generation rate per unit volume which depends on the Kane tunneling parameters. In the presence of uniform electric field, GBTBT is defined as [14] GBTBT P E B =A exp − , E0 E (3.2) where E0 = 1 V/cm, P = 2 and 2.5 for direct and indirect tunneling, respectively, A and B are the parameters for the Kane model and E is the electric field across the tunneling junction. The Kane parameters A and B for indirect tunneling are given by: Aind = g (mDOSc mDOSv )3/2 (1 + 2NTA ) DT2 A (qE0 )5/2 5/4 221/4 h5/4 mr ρTA (Eg + Mc )7/4 , (3.3) 1/2 Bind = 27/2 πmr (Eg + Mc )3/2 , 3qh (3.4) where the symbols have there conventional meaning as specified in the Ref. [1]. The conduction band density of state effective mass (mDOSC ) of the ∆4 valley is defined 1 as a geometric mean of longitudinal (ml ) and transverse (mt ) effective masses as (ml m2t ) 3 (ml and mt are nearly independent of Ge Content [13]). Valence band density of states effective 3 3 2 2 mass (mDOSV ) is defined as (mlh + mhh )(2/3) , where, mlh and mhh are the light and heavy hole effective masses, respectively. The reduced tunneling mass (mr ) is defined as by the curvature 3.4 BTBT Model Parameters 51 Table 3.1: Parameter which are used for the calculation of non-local BTBT model parameters for comre is the relaxed pressively strained SiGe at various Ge mole fraction taken from the work [1], where EG band gap. All effective masses are in the units of m0 (free electron mass). x 0 0.1 0.2 0.3 0.4 0.5 0.6 ρ (kg/m3 ) 2329 2673 3008 3332 3646 3951 4245 DT A (eV/m) 1.69 × 1010 1.64 × 1010 1.60 × 1010 1.55 × 1010 1.51 × 1010 1.46 × 1010 1.42 × 1010 T A (meV) 19 18 16.9 15.9 14.8 13.8 12.8 Egre (eV) 1.12 1.08 1.03 0.991 0.959 0.935 0.919 EgV 1 (eV) 1.12 1.08 1.03 0.991 0.959 0.935 0.919 EgV 2 (eV) 1.12 1.06 0.990 0.935 0.882 0.835 0.799 mL 0.92 0.92 0.92 0.92 0.92 0.92 0.92 mT 0.19 0.19 0.19 0.19 0.19 0.19 0.19 mDOSC 0.32 0.32 0.32 0.32 0.32 0.32 0.32 mDOS,V 1 0.52 0.23 0.21 0.20 0.18 0.16 0.15 mDOS,V 2 0.15 0.22 0.21 0.19 0.18 0.16 0.15 Table 3.2: Degeneracy factor g, electron curvature effective mass (me ), hole curvature effective mass (mV 1 / mV 2 ), and the reduced tunneling mass (mr ) used in the theoretical calculation of Aind and Bind along [001] tunneling direction in the c-SiGe. All effective mass are in the units of the free electron mass. Aind and Bind are in units of cm−3 s−1 and M V cm−1 , respectively. Note that g, me and mV 1 (V1 corresponds to HH in relaxed SiGe) are strain independent. x 0 0.1 0.2 0.3 0.4 0.5 0.6 g 8 8 8 8 8 8 8 me 0.19 0.19 0.19 0.19 0.19 0.19 0.19 mV 1 0.289 0.284 0.277 0.268 0.259 0.250 0.240 mv2 0.201 0.228 0.233 0.288 0.220 0.211 0.20 mVr 1 0.115 0.114 0.113 0.111 0.110 0.108 0.106 mVr 2 0.098 0.104 0.105 0.104 0.102 0.100 0.097 AVind1 7.41 × 1015 2.24 × 1015 2.28 × 1015 2.24 × 1015 2.31 × 1015 2.28 × 1015 2.41 × 1015 V1 Bind 24.9 22.18 19.26 17.01 14.94 13.23 11.61 AVind2 1.47 × 1015 2.40 × 1015 2.31 × 1015 2.158 × 1015 2.158 × 1015 2.08 × 1015 2.02 × 1015 V2 Bind 23 21.6 19.62 17.91 16.29 14.85 13.68 mass of conduction band (me ) and valence bands (mV 1 and mV 2 ) along the [001] tunneling me mlh direction mr = me +mlh [15]. NT A = exp(T A1/KT )−1 is the occupation number of the transverse acoustic phonon, ρ is the mass density, T A is the transverse acoustic phonon energy, and Eg is the minimum band gap [1]. All the parameters values, which are used to calculate the Kane parameters for c-SiGe are listed in Table. 3.1. The Kane BTBT tunneling parameters are tabulated in Table. 3.2, along the [001] tunneling direction. Mass density (ρ) as a function of Ge mole fraction is given by, ρ = 2.329 + 3.493x − 0.499x2 [16]. The estimated deformation potential of transverse acoustic phonons (DT A ) for Si is 2.45 ×1010 eV/m, and for Ge is 0.8 ×1010 eV/m [17]. The transverse acoustic phonon energy T A for Si is 19 meV, and for Ge is 8.6 meV. These values are extracted from [16], and for SiGe it is determined by linear interpolation [18]. We have only considered transverse acoustic phonon into account because of its high occupation number and small phonon energy [19]. Degeneracy factor is expressed as: G= 2gv gc which consists of the 3.5 Field Induced Quantum Confinement (FIQC)- Back Envelope Calculations. 52 electron spin degeneracy factor 2 and the valence and conduction band valley degeneracy. 3.5 Field Induced Quantum Confinement (FIQC)- Back Envelope Calculations. The effect of sub-band discretization due to the formation of the quantum well at the oxide/semiconductor interface is considered by back envelope calculation method [5]. In the Kane dispersion relationship, valence and conduction band are considered as continuous in the BTBT current calculations without sub-band effects. To include the sub-band effect, triangular well approximation is used and the first sub band energy is evaluated using Airys function approach as, α Esub = −a0 2 h2 Fox,α 8π 2 m∗α 13 , (3.5) where a0 =-2.381 is the first zero of the airy function, and m∗α is electron effective mass in the α valley. The force acting on the electron at the semiconductor/oxide interface Fox,α at the onset of tunneling is, s Fox,α = α 2q 2 Ns (EGα + Esub ) , s (3.6) where s is the semiconductor permitivity, and Egα is the tunneling band gap between the valence α band and the α valley of the conduction band. A cubic equation will be formed in Esub , which has two imaginary and one real roots [5]. The onset voltage of the tunneling can be written as, α α Vonset = Egα + Esub + Fox,α tox s − VF B , ox (3.7) where VF B is the flat band voltage. The graph is shifted manually by Vonset to include the effect of FIQC. 3.6 Results and Discussions 3.6 53 Results and Discussions Having described the strain effects in SiGe and Si. In this section, we discuss in detail the results that are obtained from the simulation for the different proposed device structure. The strain effects are incorporated at the source and channel. In section 3.6.1, we discuss heteroDGTFET (H-DGTFET) with HD gate. The effect of pocket implant at the source/channel junction is discussed in section 3.6.2. This is followed by the discussion of the source extended structure to find the optimum source extension length in section 3.6.3. 3.6.1 DGTFET with HD gate and c-SiGe source In this section, the device structure that is shown in Fig. 3.1a is discuss in detail. In particular, we will focus on the work function engineering, effect of different hetero dielectrics, and impact of Ge mole fraction variation in c-SiGe. 3.6.1.1 Workfunction Engineering The work function of the metal at the gate is varied and its impact on the transfer characteristics is shown in Fig. 3.3a. As the metal gate work function increases, at ON state, the energy bands are pulled upwards. This results in increased barrier width at the junction as shown in Fig. 3.3b, which decreases BTBT current as shown in Fig. 3.3a. The use of higher work function metal at the gate results in threshold voltage shift and the transfer characteristics are shifted parallel to the X-axis. As a result subthreshold slope is not affected. The OFF current in TFET is composed of BTBT current plus reverse leakage current, considering the generation and recombination in the depletion region. The OFF current remains almost constant as the metal work function increases. This is a consequence of the fact that as work function of metal increases, energy band below the gate are pulled upwards. This prevents the formation of tunneling path in the device as shown in Fig. 3.3c and thus the OFF state current is composed only of reverse leakage current. At low work function due to the formation of tunneling path in the device BTBT takes place, which increases the OFF current. The SSAV G for the proposed device structure comes out to be 64.3 mV/decade for 4.2 eV metal 3.6 Results and Discussions (a) (c) 54 (b) (d) Figure 3.3: DGTFET (L = 25 nm and tox =2 nm) based on the c-SiGe source, 0.5 Ge mole fraction, and HD gate for different values of φm . (a) Transfer characteristics at VDS = 1.0 V shown on both semi log and linear scale. Energy band diagram along a horizontal cut line near to the surface is shown at:(b) ON-state (VGS =1.5 V and VDS = 1V), and (c) OFF-state (VGS = 0.0 V and VDS = 1V). (d) Output characterisitcs for different gate voltages. work function and 0.5 Ge mole fraction in c-SiGe. The output characteristics of the devices for different gate voltage is shown in Fig. 3.3d. The drain saturation voltage (Vdsat ) is calculated as the value of drain voltage which diminishes the channel charge at the drain end and is evaluated at 95 % of the steady state value of drain current. The value of Vdsat comes out to be 0.64 V at VG =1.0 V. 3.6 Results and Discussions (a) 55 (b) Figure 3.4: DGTFET based on c-SiGe source, 0.5 Ge mole fraction and different hetero gate dielectric. (a) Energy band diagram calculated along a horizontal plane near to the surface is shown at ON state. (b) Transfer characteristics at φm =4.2 eV at VDS =1.0 V. 3.6.1.2 Effect of different hetero dielectrics The dielectric, which is used at the drain end is fixed with SiO2 and the dielectric at the source end is replaced with high-k dielectric. The dielectric that we have used are Hf O2 and ZrO2 with a dielectric constant of 22 and 29 respectively. The simulation results obtained using Hf O2 and ZrO2 are shown in Fig. 3.4 for 4.2 eV as work function of the gate metal. It is evident from Fig. 3.4a that as the value of dielectric constant increases, barrier width at the junction reduces as shown in Fig. 3.4a. The reduced barrier width is due to the improved gate coupling with high-k dielectric. The decrease of the barrier width at the junction results in increase in the BTBT current as shown in Fig. 3.4b. 3.6.1.3 Effect of Ge mole fraction in c-SiGe As the mole fraction of Ge increases in c-SiGe, compressive strain increases in the SiGe source, which reduces the band gap as shown in Fig. 3.5a. This results in reduced barrier width and the increased BTBT current as shown in Fig. 3.5b. A slight increase in the OFF current is observed due to the change in the intrinsic carrier concentration, because of reduction in bandgap with increasing Ge mole fraction. The SSAV G and ION is plotted against Ge mole fraction in Fig. 3.5c. SSAV G and ION improves with Ge mole fraction. 3.6 Results and Discussions 56 (a) (b) (c) Figure 3.5: The effect of different mole fraction of Ge(x) in DGTFET based on c-SiGe with HD gate. (a) Energy band diagram along a horizontal cut line near to the surface is shown at ON-state (VGS =1.5 V and VDS = 1 V). (b) Transfer characteristic at VDS = 1.0 V at φm =4.2 eV. (c) SSAV G and ION v/s Ge mole fraction (x). 3.6.2 DGTFET with Pocket Implant In this section, we explore the impact of pocket implant under the gate at source/body junction in a DGTFET as shown in Fig. 3.1. The pocket implant is considered to be n-type and 2 nm in length. As the pocket implant doping increases, BTBT current increases because of the reduced barrier width at the junction. The increase in ON current due to pocket implant is due to the band gap narrowing at the source/body junction. Band gap narrowing results in steeper energy band at the junctions as shown in Fig. 3.6a, which increase the BTBT current. The increase in BTBT current is more for pocket implant doping (NP OC ) greater than 1019 cm−3 as shown in Fig. 3.6b. This is because 3.6 Results and Discussions 57 (a) (b) (c) Figure 3.6: DGTFET (L = 25 nm and tox =2 nm) based on the c-SiGe source, 0.5 Ge mole fraction, HD gate and pocket implant near the source junction. (a) Energy band diagram along a horizontal cut line near the surface is shown at ON-state (VGS =1.5 V and VDS = 1 V) for different pocket implant doping (NP OC ) at φm =4.2 eV. (b) Transfer characteristics at VDS = 1.0 V is shown on both semi log and linear scale for different value of NP OC . (c) SSAV G and ION against NP OC doping. as the value of NP OC reaches the density of states value, Fermi level moves into the conduction band. This reduces the effective band gap and results in a discontinuity in the conduction band, which appeared in the form of kink in Fig. 3.6a. The OFF state current rises as the value of NP OC increases. The increase in OFF state current is entirely attributed due to the SRH field enhanced recombination. SRH recombination lifetime reduces because of high electric field, which causes electron hole pair generation before BTBT tunneling and increases the OFF state current by a large amount. There are three main contributor to SRH recombination, which changes the minority carrier life time and hence the generation and recombination of carriers. The minority carrier lifetime is modeled as product 3.6 Results and Discussions (a) 58 (b) Figure 3.7: Transfer characteristics of DGTFET with extended c-SiGe source and 0.5 Ge mole fractions for different encroachment length (nm) inside the gate. As the source is extended inside the gate, tunneling is delayed due to the increase in tunnel barrier width because of the depletion of the extended part of the source. (b) Subthreshold slope and ON-current for different extension length inside the gate. of doping dependence, field dependent and temperature dependent lifetime. The SRH doping dependence has been taken into account by Scharfater model, which models the carrier lifetime as a function of doping. The enhanced electric field reduces SRH recombination lifetime in regions of strong electric field and it is considered by SRH field enhanced model. In Fig. 3.6c, we have shown the variation of SSAV G and ION with pocket implant doping (NP OC ). The SSAV G and ON current improves with pocket implant doping. The effect of FIQC can be considered by manually shifting the curve by Vonset . The calculation of Vonset is discussed in section 3.5. The decrease in band gap due to band gap narrowing for NP OC =1e+20 cm−3 comes out to be 0.032 eV (extracted from simulations). The value of onset voltage for NP OC =1e+20 cm−3 at Eg =1.12-0.032 eV comes out to be 0.956 V and the value of Fox,α comes out to be 3.47 MeV/cm. Without pocket implant the value of onset voltage comes out to be 0.988 V. 3.6.3 DGTFET with source extended structure In this section, the source extended structure, which is shown in Fig. 3.1c is discuss. The transfer characteristics of the device is shown in Fig. 3.7a. It is expected that the extended p++ doping underneath the gate results in depletion in the extended part of source leading to 3.6 Results and Discussions (a) 59 (b) Figure 3.8: Energy band diagram along a horizontal cut near to the surface calculated at: (a) ON state (VGS =1.5 V and VDS =1 V). The tunnel barrier width increases with the extension length and 2nm extension gives the best BTBT rate.(b) OFF state (VGS =0 V and VDS =1 V). No tunneling path is formed for any of the structures in the OFF state as shown by dotted line. a widening source-channel barrier width. This concomitant lower BTBT rate, and increasing SS with increasing source extension. The extracted subthreshold slope (SS) and ION current variation for 0, 2, 4 and 6 nm extension comes out to be 67.2, 65.7, 78.3 and 71.6 mV/decades respectively, as shown in Fig. 3.7b. It is also observed that the ON-current is maximum for the 2 nm extension as shown in Fig. 3.7b. The reason for this behavior can be explained as follows: tunneling path is always shorter for the 0 nm extension as shown in Fig. 3.8a. However, the tunneling, which is from c-SiGe valence band to t-Si conduction band, mostly occurs in the Si region. On the other hand, for the 2 nm extension, at larger VG , the whole tunneling path lies completely inside SiGe, which results in higher BTBT rate due to lower reduced mass. In the OFF-state, the BTBT current in the TFET is comparable or less than the reverse leakage diode current, which is dependent on the generation current inside the depletion region. It has been observed from Fig. 3.8b, in the OFF-state, the source valence band and the channel conduction band do not overlap and there is no tunneling path at the junctions. Hence OFF current is limited by the SRH recombination leakage current. This generation rate is higher in SiGe than in Si due to its lower band gap and higher intrinsic carrier concentration. Extending the SiGe under the gate also results in depletion in the overlap region, resulting in higher SRH 3.6 Results and Discussions (a) 60 (b) Figure 3.9: (a) Transfer characteristics of DGTFET with extended c-SiGe source, relaxed SiGe source (r-SiGe) (Ge mole fraction = 0.5) and all Si TFET. (b) The effect of FIQC in tunneling showing the current components and the shift in the tunneling. leakage currents, which gives the saddle points in the ID -VG at low VG . 3.6.3.1 Si, r-SiGe (relaxed SiGe), and c-SiGe extended source TFET The comparison between Si-TFET, TFET with r-SiGe (Relaxed SiGe) source and TFET with c-SiGe source has been made in Fig. 3.10a. It can be concluded that the compressive strain at the source increases the ON-current, because of reduction in bandgap. The two tunneling path, which are formed due to the splitting of valence band increase the current by large amount as compared to relaxed SiGe source. The SSAV G comes out to be 83.7, 79.7 and 65.7 mV/decade for all Si, r-SiGe and c-SiGe extended source TFET. Significant improvement in both the ONcurrent and the SSAV G can be seen using compressive strain at the source. The effect of FIQC is considered in the onset voltage. The graph is shifted manually by Vonset , however the current before the onset of tunneling will consists of reverse saturation current. We merge these two current at the onset voltage to get the characteristics as shown in Fig. 3.10b. The decrease in band gap due to band gap narrowing for high doping comes out to be 0.06 eV (extracted from simulations). The value of onset voltage in tunneling at Eg=0.756-0.06 eV comes out to be 0.732 V and the value of Fox comes out to be 3.1 MeV/cm. 3.7 Summary 61 (a) (b) Figure 3.10: Channel length scaling of DGTFET (a) with gate aligned structure, and (b) Source extended structure. 3.6.4 Channel length scaling We have also performed simulation for different channel lengths for gate aligned and source extended structure. The channel length scaling should not affect the BTBT current. This is due to the fact that BTBT takes place only few nm along side source/body junction. As channel length is scaled, electrostatics of the system is changed slightly, which affects the OFF current. These results are also verified from the simulations and are shown in Fig. 3.10. 3.7 Summary This work has emphasized upon the compressive strain effect on SiGe. We have investigated different device structure and have observed that the compressive strain in SiGe, improves the SS and ON current as compared to r-SiGe and Si TFETs. As a result of the compressive strain, splitting of valence and conduction band takes place and the effective band gap reduces. Additionally, from pocket implant study it can be concluded that its not a effective method to increase the ON current due to very large increase in OFF current. The hetero dielectric gate results in simultaneous decrease in the OFF current and increase in the ON current, thus good performance can be achieved. The source extended structure are also scrutinized to improve 3.7 Summary 62 the TFET performance. It is observed that 2 nm source extension inside the gate results in the best ON current and subthreshold slope. Different extensions length inside the gate degrades the sub-threshold slope, but the increase in ON state current is large. Bibliography [1] K.-H. Kao, A. S. Verhulst, R. Rooyackers, B. Douhard, J. Delmotte, H. Bender, O. Richard, W. Vandervorst, E. Simoen, A. Hikavyy, R. Loo, K. Arstila, N. Collaert, A. Thean, M. M. Heyns, and K. D. Meyer, “Compressively strained SiGe band-to-band tunneling model calibration based on p-i-n diodes and prospect of strained SiGe tunneling field-effect transistors,” Journal of Applied Physics, vol. 116, no. 21, p. 214506, 2014. [2] N. D. Chien, L. T. Vinh, N. V. Kien, J. K. Hsia, T. S. Kang, and C. H. Shih, “Proper determination of tunnel model parameters for indirect band-to-band tunneling in compressively strained SiGe TFETs,” in International Symposium on Next-Generation Electronics, Feb 2013, pp. 67–70. [3] L. Yang, J. R. Watling, R. C. W. Wilkins, M. Borii, J. R. Barker, A. Asenov, and S. Roy, “Si/SiGe heterostructure parameters for device simulations,” Semiconductor Science and Technology, vol. 19, no. 10, p. 1174, 2004. [4] K. W. Ang, J. Lin, C. H. Tung, N. Balasubramanian, G. S. Samudra, and Y. C. Yeo, “Strained-MOSFET With Embedded Source/Drain Stressors and Strain-Transfer Structure (STS) for Enhanced Transistor Performance,” IEEE Transactions on Electron Devices, vol. 55, no. 3, pp. 850–857, Mar 2008. [5] W. G. Vandenberghe, B. Sore, W. Magnus, G. Groeseneken, and M. V. Fischetti, “Impact of field-induced quantum confinement in tunneling field-effect devices,” Applied Physics Letters, vol. 98, no. 14, p. 143503, 2011. 63 BIBLIOGRAPHY 64 [6] A. Murthy, R. S. Chau, T. Ghani, and K. R. Mistry, “Semiconductor transistor having a stressed channel,” Sept 2003, US Patent 66 211 31. [7] M. Schmidt, L. Knoll, S. Richter, A. Schafer, J. M. Hartmann, Q. T. Zhao, and S. Mantl, “Si/SiGe hetero-structure tunneling field effect transistors with in-situ doped SiGe source,” in International Conference on Ultimate Integration on Silicon (ULIS), Mar 2012, pp. 191–194. [8] Z. Cheng, J. Jung, M. L. Lee, A. J. Pitera, J. L. Hoyt, D. A. Antoniadis, and E. A. Fitzgerald, “Hole mobility enhancement in strained-Si/strained-SiGe heterostructure pMOSFETs fabricated on SiGe-on-insulator (SGOI),” Semiconductor Science and Technology, vol. 19, no. 5, p. L48, 2004. [9] G. Lee, J.-S. Jang, and W. Y. Choi, “Dual-dielectric-constant spacer hetero-gate-dielectric tunneling field-effect transistors,” Semiconductor Science and Technology, vol. 28, no. 5, p. 052001, 2013. [10] A. S. Verhulst, W. G. Vandenberghe, K. Maex, and G. Groeseneken, “Tunnel field-effect transistor without gate-drain overlap,” Applied Physics Letters, vol. 91, no. 5, p. 053102, 2007. [11] D. J. Robbins, L. T. Canham, S. J. Barnett, A. D. Pitt, and P. Calcott, “Near band bgap photoluminescence from pseudomorphic SiGe single layers on silicon,” Journal of Applied Physics, vol. 71, no. 3, pp. 1407–1414, 1992. [12] T. Manku and A. Nathan, “Energy-band structure for strained p-type Si1−x Gex ,” Phys. Rev. B, vol. 43, pp. 12 634–12 637, May 1991. [13] M. V. Fischetti and S. E. Laux, “Band structure, deformation potentials, and carrier mobility in strained Si, Ge, and SiGe alloys,” Journal of Applied Physics, vol. 80, no. 4, pp. 2234–2252, 1996. [14] E. O. Kane, “Theory of Tunneling,” Journal of Applied Physics, vol. 32, no. 1, pp. 83–91, 1961. BIBLIOGRAPHY 65 [15] J. Spitzer, K. Thonke, R. Sauer, H. Kibbel, H. Herzog, and E. Kasper, “Direct observation of band edge luminescence and alloy luminescence from ultrametastable silicon germanium alloy layers,” Applied Physics Letters, vol. 60, no. 14, pp. 1729–1731, 1992. [16] C. Rivas, “Full bandstructure modeling of indirect phonon assisted transport in tunnel devices with silicon contacts,” Sept 2003, Ph.D. dissertation, Univ. Texas, Dallas. [17] M. V. Fischetti, T. P. O’Regan, S. Narayanan, C. Sachs, S. Jin, J. Kim, and Y. Zhang, “Theoretical Study of Some Physical Aspects of Electronic Transport in nMOSFETs at the 10-nm Gate-Length,” IEEE Transactions on Electron Devices, vol. 54, no. 9, pp. 2116– 2136, Sept 2007. [18] R. A. Logan, J. M. Rowell, and F. A. Trumbore, “Phonon Spectra of Ge-Si Alloys,” Phys. Rev., vol. 136, pp. A1751–A1755, Dec 1964. [19] A. Rahman, M. S. Lundstrom, and A. W. Ghosh, “Generalized effective-mass approach for n-type metal-oxide-semiconductor field-effect transistors on arbitrarily oriented wafers,” Journal of Applied Physics, vol. 97, no. 5, p. 053702, 2005. Chapter 4 Band to Band Tunneling in Gamma Valley For Ge Source TFET: Thickness scaling 4.1 Introduction The previous chapters presented the Si, r-SiGe and c-SiGe source TFET with the different device architecture to improve the TFET performance. Further, we have discussed the workfunction and strain engineering for DGTFET as well. In this chapter, we have utilized Ge source in order to boost the current in DGTFET. Germanium (Ge) is a widely used source material because of its low band gap [1–3] and compatibility with Silicon (Si) [4]. The direct and indirect valley in Germanium (Ge) are separated by a very small offset, which opens up the prospect of direct tunneling in the Γ valley of Ge tunnel field effect transistor (TFET). We have investigated a new DGTFET design with Ge source TFET, which is based on point to point lateral tunneling. Most of the theoretical and simulation studies on Ge TFET include only the indirect tunneling with the default TCAD parameters [5]. However, the direct valley of Ge contributes significantly to various processes in Ge, including BTBT current. Moreover, Ge can be easily made to be a direct bandgap semiconductor by high tensile strain (1.8 % ) [6] or by the incorporation of the larger atom Sn in the Ge lattice [7]. At room temperature, the direct valley of Ge has been been shown (Ge on Si) to give rise to direct band gap electroluminescence [8], and has also been utilized for photonics application [9, 10]. The dominance of direct tunneling in Ge based diodes [11–13] has also been reported earlier. 66 4.2 Device Structure and Simulation Details 67 Figure 4.1: Schematic of the double gate TFET with extended Ge source and hetero-dielectric-gate (HG). We consider a device of channel length 40 nm, n-type drain doping of 1018 cm−3 , p-type source doping of 1020 cm−3 , and body is low n-type doped (1016 cm−3 ), HG gate dielectric with 2 nm thickness. Source is extended inside the gate by 2 nm. Work function of metal at the gate is Φm = 4.2 eV In Si, the difference between direct and indirect valley is 2.3 eV, which prohibits the prospects of direct tunneling in the Si based TFET’s [14]. The corresponding valley offset in bulk Ge is 0.14 eV [15], which increases the probability of direct tunneling in the Ge based TFETs. It has already been shown that thickness scaling changes the bandstructure parameters significantly in various devices [16–18]. Therefore, in this chapter, we explore the impact of Ge and Si thickness scaling on bandstructure parameters (band gap and effective mass), their relative band alignment, Kane’s BTBT parameters and the BTBT tunneling current in Ge source lateral TFET. The chapter is organized as follows: Device structure along with the band structure results, obtained from DFT calculations, and the simulations models that are used to calculate the BTBT current are discussed in section 4.2. Section 4.3 explains the theoretical calculation of BTBT model parameters for ultra thin Ge and Si thickness. This is followed by results and discussion in section 4.4. Finally we summarize our findings in section 4.5. 4.2 4.2.1 Device Structure and Simulation Details Device Structure The device structure under study is shown in Fig. 4.1 with source made of Ge to increase the BTBT probability. To reduce the ambipolar and OFF currents, we have used hetero-dielectricgate [19]. Since Ge has better BTBT parameters as compared to Si, to enhance the BTBT current, the Ge source is extended inside the gate by 2 nm to confine tunneling inside Ge [20, 21]. The study comprises of ultra thin (≤ 10nm) source lateral TFETs. In BTBT current calcula- 4.2 Device Structure and Simulation Details 68 tions, Kane’s BTBT tunneling parameters for ultra thin Ge and Si slabs is calculated by considering bandstructure parameters, which are obtained through first principle calculations after incorporating geometrical confinement. This results in the accurate evaluation of the theoretical BTBT parameters. To account the Si thickness scaling effect on BTBT current, we have considered the confined bandstructure parameter for Si as well. 4.2.2 Atomistic Simulation The thickness dependent Ge and Si band gap values are obtained from density functional calculations, by simulating the confined Ge and Si super cell using QuantumWise Atomistic Simulation package [22]. Electron-electron interactions and correlation effects have been considered under meta Generalized Gradient approximation (mGGA) and a vacuum region of 1.5 nm is added to both sides of Ge and Si super cell to prevent the interactions between the spurious replica images. The dangling bond on both sides of the Ge and Si layer are passivated by hydrogen atoms, to prevent mid-gap states. Following Ref. [16], we have assumed that the structural properties like inter-layer distance, bond-length etc. remain unaffected due to transverse confinement and used the bulk lattice constant of Ge and Si. We have calculated the thickness dependent electronic dispersion at (100) Ge and Si surface orientations [16]. As a first step, we have calibrated bulk Ge and Si bandstructure with the existing bulk band gap values using MGGA. Then the same simulation parameters for MGGA are used in all other calculations. The confined band structure of Ge at 4 nm thickness is shown in Fig. 4.2a. Confinement increases the band gap of indirect valley from 0.66 eV bulk to 0.82 eV at 4 nm thickness. Further, we note that the direct valley is not properly captured in ATK simulation (even for bulk Ge) due to the very small direct valley offset. To overcome this and for consistency with the reported bulk values, a constant offset at direct valley is added for all thickness [24, 25]. For Si, the confined band structure at 6 nm thickness is shown in Fig. 4.2b. As expected the confinement increases the band gap in Si from 1.12 eV in bulk to 1.176 eV in indirect valley. 4.2 Device Structure and Simulation Details (a) 69 (b) Figure 4.2: The confined bandstructure for: (a) Ge 4 nm film and (b) Si 6 nm thin film. The confinement increases the band gap in both cases. Note that Si becomes direct band gap semiconductor at reduced thickness [23]. 4.2.3 TCAD Simulation Models For the calculation of BTBT current, Synopsys Sentaurus TCAD simulator has been used, where we have included all the relevant physical effects. BTBT current was calculated by using non-local BTBT tunneling model [26]. Note that the default BTBT model parameter for Ge in Sentaurus is for indirect tunneling. To overcome this deficiency, we have theoretically calculated the parameters for both direct and indirect tunneling for different Ge thickness at (100) orientations as tabulated in Table 4.2, and used those parameters to calculate BTBT current. The thickness dependent Si BTBT parameter are also evaluated from the DFT results and are tabulated in Table 4.3. To include non-local electric field effects hydrodynamic transport is considered and to account for the very high doping in source, band gap narrowing and Fermi Dirac statistics are used. Note that the geometrical quantum effects in BTBT tunneling have been explicitly considered by theoretically calculating BTBT parameters after considering the thickness dependent geometrical confinement effects from atomistic simulations. The default band-gap is changed with the effective thickness dependent band gap, which we have obtained from DFT simulations. The charge distribution due to the charge centroid shift at oxide/semiconductor interface is captured using MLDA model. Note that ideally, there will also be stress at Si/Ge interface. However, we have not included it in our calculations due to the 4.3 BTBT Parameter and Current Calculations 70 huge computational resources required. 4.3 BTBT Parameter and Current Calculations The BTBT current is defined as the integral of the generation rate over the entire device region as described in section 3.4 [26]. 4.3.1 Effective Masses The curvature masses and density of state (DOS) effective masses for indirect and direct valley are needed to evaluate the Kane BTBT parameters in Sentaurus. The effective masses are extracted from the band structure results that we have obtained from DFT simulations, by using parabolic fitting. The density of states (DOS) mass in conduction band is defined as mDOSc = (ml m2t )1/3 where ml (mt ) denotes the longitudinal (transverse) effective mass. Similarly, the 2/3 3/2 3/2 DOS mass in valence band is mDOSv = mlh + mhh , where mhh (mlh ) is the heavy hole (light hole) effective mass. The reduced tunneling mass mr is defined by the curvature mass of conduction band (me ) and valence bands corresponding to the light hole along the tunneling direction: mr = me mlh /(me + mlh ) [27]. 4.3.2 Kane Parameters A and B The Kane parameters A and B for direct and indirect tunneling are given by: Adir = 9 π2 1/2 qπmr (qE0 )2 1/2 , 9h2 EgΓ + Mc 1/2 Bdir = Aind = π 2 mr EgΓ + Mc qh 3/2 , g (mDOSc mDOSv )3/2 (1 + 2NTA ) DT2 A (qE0 )5/2 5/4 (4.1) 221/4 h5/4 mr ρTA (Eg + Mc )7/4 (4.2) , (4.3) 4.3 BTBT Parameter and Current Calculations 71 Table 4.1: Calculated density of states for indirect and direct valley. Nv , NcL and NcΓ is the DOS at valence band, conduction band of indirect and direct valley, respectively. χL and χΓ is the electron affinity at indirect and direct valley. ∆c(L−Γ) and ∆c(Γ−Γ ) are the conduction band offset at Si/Ge hetero junction considering the indirect and direct valley of Si. T hickness Bulk (≥ 10nm) 8nm 6nm 4nm Nv 5.1e18 5.35e18 5.83e18 7e18 Germanium NcL NcΓ 1.02e19 2.9e17 1.03e19 2.9e17 1.09e19 3.27e17 1.23e19 4.02e17 χL 4.1 4.07 4.03 3.94 χΓ 3.96 3.95 3.91 3.81 NcΓ 3e18 3.01e18 3.04e18 3.11e18 Silicon NcL χL 4.24e19 4.05 4.33e19 4.04 4.54e19 4.03 4.93e19 3.99 χΓ 4.048 4.037 3.99 3.94 Band Offset at Si/Ge ∆c(L−Γ) ∆c(Γ−Γ ) 0.09 0.088 0.09 0.087 0.11 0.08 0.18 0.13 1/2 Bind = 27/2 πmr (Eg + Mc )3/2 , 3qh (4.4) where g is the degeneracy factor, DTA , TA and NTA is the deformation potential, energy and occupation number of the transverse acoustic phonon respectively. Mc is the conduction band offset at the tunneling junction. At low gate voltage, tunneling is taking place from Ge valence band to Si conduction band. For this tunneling, Mc is calculated from the difference in electron affinity of Si and Ge. All other parameters have their conventional meaning. In the calculation of the Kane’s parameter, DTA = 0.8 × 1010 eV/m and TA is 8.6 meV for Ge [28, 29], while for Si it is DTA = 2.45 × 1010 eV/m and TA is 19 meV. 4.3.3 Effect of DOS in the effective electric field The impact of DOS on the BTBT current is two fold: a) as evident from Eqs. 4.1-4.3, the Kane’s tunneling parameters A and B have DOS dependence in terms of the DOS effective mass, b) The DOS of the individual valleys has to be included in the carrier distribution as well which is an input to the 2-D Poisson equation (with quasi static approximation) used to obtain the surface potential profile in the device. The impact of the direct valley DOS on the internal electric filed has been neglected in earlier works [30], and is also a shortcoming of Sentaurus in which the default parameter file for Ge uses the indirect valley DOS for electrostatic calculations. The effective DOS in the conduction and valence band for the direct and indirect valley can 4.3 BTBT Parameter and Current Calculations 72 Table 4.2: Bandstructure dependent parameters obtained from atomistic simulation for Ge (100) slab. L and E Γ are the band gap of indirect and direct valleys in eV . Theoretically calculated BTBT Here, EG G model parameters for direct and indirect tunneling for Ge (100) slab at different thickness. A and B are in units of cm−3 s−1 and M V cm−1 , respectively. Germanium Thickness bulk ≥ 10nm 8nm 6nm 4nm Band Structure Parameters ml mt EGΓ mΓ 0.66 1.6 0.08 0.8 0.04 0.69 1.62 0.08 0.81 0.044 0.73 1.61 0.085 0.85 0.047 0.82 1.65 0.096 0.95 0.054 EGL Direct BTBT ADir BDir 20 1.465 × 10 6.03 1.46 × 1020 6.14 1.45 × 1020 6.71 1.42 × 1020 8.19 Indirect BTBT Aind Bind 15 2.23 × 10 4.98 2.05 × 1015 5.39 1.89 × 1015 6 1.64 × 1015 7.47 be calculated as: Nc Nv 3/2 2πmDOSc kT = 2M , h2 3/2 2πmDOSv kT = 2 . h2 2/3 (4.5) (4.6) where M is the number of equivalent valleys in the conduction band. The effective mass, and consequently the DOS for both the valleys also change with thickness scaling. So, we have theoretically calculated the DOS for direct and indirect valley of Ge and Si and is shown in Table 4.1. 4.3.4 Effect of band alignment at Si/Ge junction As we scale down in thickness, the band alignment at the Si/Ge hetero junction changes significantly which affects the BTBT current. Due to the difference in electron affinity and band gap of Si and Ge, a discontinuity is present at both valence and conduction band at source/channel junction. Jaro’s theory predict a valence band offset of 4Ev,av = 0.48 eV for a bulk Ge/Si hetero interface [7], which is in reasonable agreement with the experimentally reported value of 0.5 eV [7]. The electron affinity also changes with thickness scaling. It is calculated as per Ref. [39], and tabulated in Table 4.1. The conduction band offset at Si/Ge interface for various thickness is evaluated and is tabulated in Table 4.1 for both, direct and indirect valley of Si. Table 4.2 contains all the bandstructure parameter values obtained from the atomistic sim- 4.4 Results and Discussion 73 L Table 4.3: Bandstructure parameters obtained from atomistic simulation for Si (100) slab. Here, EG Γ and EG are the band gap of indirect and direct valleys in eV . Theoretically calculated BTBT model parameters for direct and indirect tunneling for Si (100) slab at different thickness. A and B are in units of cm−3 s−1 and M V cm−1 , respectively. Silicon Thickness bulk ≥ 10nm 8nm 6nm 4nm Band Structure Parameters ml mt EGΓ mΓ 1.12 1.4 0.354 1.105 0.206 1.133 1.402 0.361 1.114 0.206 1.176 1.405 0.378 1.138 0.208 1.225 1.416 0.409 1.162 0.211 EGL Direct BTBT ADir BDir 2.44 × 1020 22.3 2.431 × 1020 22.5 2.392 × 1020 23.8 2.31 × 1020 26.6 Indirect BTBT Aind Bind 4.68 × 1015 29.85 4.62 × 1015 30.6 4.53 × 1015 32.2 4.18 × 1015 36.5 ulation, that are used to calculate BTBT model parameter for Ge and the other parameters are taken from Ref. [30]. The Kane’s parameter A and B for Ge are evaluated by considering all the above mentioned effect and are also tabulated in Table 4.2. As Ge thickness is scaled, the Kane parameter A decreases while B increases. For BTBT parameters calculated from DFT after incorporating the above mentioned effects, the value of Bdir for Ge at 8 nm thickness increases from 6.14 MV/cm to 8.19 MV/cm at 4 nm thickness, which should degrade the current. As a consistency check we note that for Bulk Ge (100), our Bdir = 6.03 MV/cm, is more or less consistent with earlier reported values of 5.3 MV/cm and 5.7 MV/cm in Refs. [13] and [12]. Note that the default value of Kane’s parameters used for Ge in Sentaurus TCAD are A = 9.1 × 1016 cm−3 s−1 and B = 4.9 MV/cm [29, 31], which grossly overestimate the current at reduced thickness. As expected the similar trend is followed for Si thickness dependent BTBT parameters. The value of A decreases, while B increases. The thickness dependent Si BTBT parameters are shown in Table 4.3. 4.4 Results and Discussion Having described the calculations of the BTBT parameters for thickness scaled Ge and Si, in this section, we discuss the impact of thickness scaling on the BTBT current by using the DFT calculated BTBT parameters. The band alignment will change significantly with thickness scaling due to the confinement effect in Si and Ge. To investigate the impact of band alignment with thickness, we have separately discussed the Ge to Si and Ge to Ge tunneling in the following 4.4 Results and Discussion (a) 74 (b) Figure 4.3: (a) Thickness scaling of total BTBT current considering direct and indirect valley of Si (VDS = 0.7 V). Subsurface BTBT improves the current in subthreshold region with thickness scaling of Ge films due to the improved gate coupling. (b) Variation of ON current with Ge thickness for a constant gate overdrive voltage of 0.8 V. Ge film thickness scaling results in decrease of ION (due to increase in band gap) and improvement in average subthreshold slope (because of the improved gate coupling). subsections. 4.4.1 Ge to Si: Band alignment and BTBT DFT results of confined Si shows that Si becomes a direct band gap material at scaled thickness [23], see Fig. 4.2b. Thus, for Si we have considered two tunneling path, both originating from valence band of Ge to direct and indirect valley of Si, with the direct and indirect tunneling BTBT parameters, respectively. We find that the direct tunneling in Si dominates over indirect tunneling, due to the better values of BTBT parameter as shown in Table 4.3. The drain current calculated for ultra thin Ge devices considering the direct and indirect valley of Si is shown in Fig. 4.3a. Note that there is a change in the trend of drain current with thickness, observed at the cross over point which appears at approximately VGS = 0.44 V (marked explicitly in the figure). This can be explained using the electrostatic effects on tunneling. The electrostatics of the system improves significantly with thickness scaling, due to the better gate coupling. This results in a substantial contribution to the total current from the sub-surface BTBT and also results in improved average subthreshold slope (SSAVG ) with reduced IOFF . The SSAVG is computed as the inverse of slope in the transfer characteristics 4.4 Results and Discussion (a) 75 (b) Figure 4.4: Energy band diagram along a horizontal cut line at distance of 1Å from the surface (at VDS = 0.7 V) for: (a) different values of gate voltage at 4 nm thin film and (b) at OFF state (VGS = 0V) for different ultra thin thickness. In (a) as gate voltage increases tunneling shifts from Ge-Si junction to Ge. In (b) as thickness is scaled, the barrier gets wider due to the confinement effect on Si and Ge. The voltage range over which tunneling takes place from Ge to Si increases with thickness scaling, because of the change in band alignment. between VT (VGS at IDS of 1e-7 A/µm) and VOFF . The ON current degrades remarkably with thickness scaling, as shown in Fig. 4.3b. Band alignment at the Si/Ge hetero junction will change as a result of the confinement effect on Si and Ge. Variation of the tunneling path with the applied gate voltage is shown in Fig. 4.4a, for 4 nm ultra thin film. It can be clearly seen from the figure that as gate voltage increases the tunneling path shifts from Ge-Si junction towards Ge. Similar trend can be expected for other thickness as well. The variation of the tunnel path for different thickness at OFF state is shown in Fig. 4.4b. The confinement effects results in increase in the band gap of Si and Ge as thickness is scaled. As a result the bands move up and the tunneling barrier gets wider, which consequently reduced the BTBT current from Ge to Si. Further, as the thickness is scaled the voltage range of tunneling from Ge-Si junction increases, which is due to the change in the band alignment. The OFF state current in TFET’s is mainly composed of BTBT current and diode leakage current, considering the generation and recombination of carriers in the depletion region (Shockley-Read-Hall or SRH currents). There are three main contributors to SRH recombination: (a) doping dependence, (b) field enhancement and (c) temperature dependence, which 4.4 Results and Discussion 76 Figure 4.5: Transfer characteristics of DGTFET with extended Ge source, where the total current is broken down into its constituent components of BTBT and SRH current at 4 nm thickness. At OFF state (VGS = 0V), the total current consists of both SRH and BTBT current. changes the minority carrier life time and hence the generation and recombination of the carriers. Smaller band gap results in increased carrier generation inside the depletion region, because of SRH electric field enhancement. Thus, SRH current decreases as thickness is reduced because of increased band gap. At OFF state, there is tunneling path from Ge to Si, however, the tunneling path associated with this process is large, as shown in Fig. 4.4b. Consequently, the total current at low gate voltage consist of both SRH and BTBT current. To further investigate the BTBT and SRH component in total current, we have broken down the total current into BTBT and SRH leakage currents, as shown in Fig. 4.5 for 4 nm thickness. At OFF state (VGS = 0V), IBTBT and ISRH constitute the total current. The tunneling current in the device at negative VGS is not affected, as the bands are pulled upward without affecting the electrostatics at the junction (see Fig. 4.4a). Thus the total current for negative VGS at 4 nm thickness is predominantly SRH currents while at higher thickness the OFF current comprise of both SRH and BTBT. The OFF state current decreases as thickness is scaled on account of the increased band gap with reduced thickness. 4.4.2 Ge to Ge:Band alignment and BTBT As the gate voltage increases tunneling path shifts from Ge-Si junction to Ge source, due to the 2 nm Ge source overlap. To analyze the tunneling in Ge, we have shown BTBT carrier 4.4 Results and Discussion 77 Figure 4.6: BTBT generation rate contour for different VGS at VDS = 0.7 V for different thickness. At VGS = 0.2 V, the BTBT generation rate is higher for 6 nm thickness, which is due to the simultaneous effect of the improved electrostatics of the system and better BTBT parameter value. At low VGS , there is tunneling from Ge to Si. For VGS = 0.6 V, the BTBT generation rate of 8 nm thin film is higher as compared to the other thickness. This is due to the improved gate coupling with VGS . generation rate contour in the device in Fig. 4.6, for different thickness at different VGS . Two bias points are considered: one at low gate voltage where mostly the tunneling is taking place from Ge to Si, and the other at high gate voltage where the tunneling is in Ge only. Now, at low gate voltage (VGS = 0.2 V), the BTBT generation rate is maximum for 6 nm thickness as shown in Fig. 4.6. This is due to the combined effect of improved gate coupling, which changes the electrostatics of the system and better BTBT parameter values than at 4 nm thickness. But, the total current is higher for 4 nm thickness at VGS = 0.2 V, due to the large effective tunneling area. As VGS increases, the tunneling gets confined in the Ge only, i.e. now the tunneling is from valence band of Ge to conduction band of Ge. This tunneling can be both direct and indirect. However, it has already been shown in various theoretical and experimental work that direct tunneling dominates in Ge [12, 14, 32]. The improved gate control and increased effective tunneling area at higher VGS in 8 nm film results in improved carrier generation, as shown in Fig. 4.6 (for VGS = 0.6 V). This results in higher current for 8 nm thickness. 4.5 Summary 4.5 78 Summary To summarize, we explore the thickness dependent BTBT current for 2 nm extended Ge source TFET. We use DFT calculations to calculate the thickness dependent bandstructure parameters such as the bandgap and effective masses etc. for Ge and Si slabs. These are then used to obtain accurate Kane’s tunneling parameters to estimate the tunneling current. The band alignment at Si/Ge junction changes significantly with thickness scaling. It is found that at reduced thickness, 2 nm overlap length is insufficient to confine the tunneling in Ge. In ultra thin devices, the electrostatic effects also play an important role in BTBT due to enhanced gate coupling, and this enhances the subsurface BTBT current significantly. The increased contribution of the subsurface BTBT current to the total drain current in ultra-thin Ge TFET results in significantly improved subthreshold region characteristics. The calculated ION at over drive voltage of 0.8 V varies from 383 µA/µm to 178 µA/µm, as we move down from 10 nm to 4 nm thickness respectively, for VDS = 0.7 V. While, the SSAVG improves from 43.5 mV/decade (10 nm) to 33.4 mV/decade (4 nm). Bibliography [1] O. M. Nayfeh, C. N. Chleirigh, J. Hennessy, L. Gomez, J. L. Hoyt, and D. A. Antoniadis, “Design of Tunneling Field-Effect Transistors Using Strained-Silicon/StrainedGermanium Type-II Staggered Heterojunctions,” IEEE Electron Device Letters, vol. 29, no. 9, pp. 1074–1077, Sept 2008. [2] S. H. Kim, H. Kam, C. Hu, and T. J. K. Liu, “Germanium-source tunnel field effect transistors with record high ION /IOFF ,” in Symposium on VLSI Technology, 2009, pp. 178–179. [3] G. Han, P. Guo, Y. Yang, L. Fan, Y. S. Yee, C. Zhan, and Y.-C. Yeo, “Source Engineering for Tunnel Field-Effect Transistor: Elevated Source with Vertical SiliconGermanium/Germanium Heterostructure,” Japanese Journal of Applied Physics, vol. 50, pp. 04DJ07–1–04DJ07–4, 2011. [4] G. Han, P. Guo, Y. Yang, C. Zhan, Q. Zhou, and Y.-C. Yeo, “Silicon-based Tunneling Field-Effect Transistor with Elevated Germanium Source formed on (110) Silicon Substrate,” Applied Physics Letters, vol. 98, no. 15, pp. 153 502–1–153 502–3, 2011. [5] “Sentaurus Device, Synopsys, Version D-2010.03,” 2016. [Online]. Available: https://www.synopsys.com/silicon/tcad/device-simulation/sentaurus-device.html [6] X. Sun, J. Liu, L. C. Kimerling, and J. Michel, “Toward a Germanium Laser for Integrated Silicon Photonics,” IEEE Journal of Selected Topics in Quantum Electronics, vol. 16, no. 1, pp. 124–131, Jan 2010. 79 BIBLIOGRAPHY 80 [7] Sun, G., Soref, and R. A. Cheng, “Design of an electrically pumped SiGeSn/GeSn/SiGeSn double-heterostructure midinfrared laser,” Journal of Applied Physics, vol. 108, no. 3, 2010. [8] S.-L. Cheng, J. Lu, G. Shambat, H.-Y. Yu, K. Saraswat, J. Vuckovic, and Y. Nishi, “Room temperature 1.6 µm electroluminescence from Ge light emitting diode on Si substrate,” Opt. Express, vol. 17, no. 12, pp. 10 019–10 024, Jun 2009. [9] J. Liu, R. Camacho-Aguilera, J. T. Bessette, X. Sun, X. Wang, Y. Cai, L. C. Kimerling, and J. Michel, “Ge-on-Si optoelectronics,” Thin Solid Films, vol. 520, no. 8, pp. 3354 – 3360, 2012. [10] R. E. Camacho-Aguilera, Y. Cai, N. Patel, J. T. Bessette, M. Romagnoli, L. C. Kimerling, and J. Michel, “An electrically pumped germanium laser,” Opt. Express, vol. 20, no. 10, pp. 11 316–11 320, May 2012. [11] A. G. Chynoweth and W. L. F. et al., “Internal Field Emission at Narrow Silicon and Germanium p-n Junctions,” Phys. Rev., vol. 118, no. 2, pp. 425–434, Apr 1960. [12] M. S. Tyagi, “Determination of Effective Mass and the Pair Production Energy for Electrons in Germanium from Zener Diode Characteristics,” Japanese Journal of Applied Physics, vol. 12, no. 1, pp. 106–108, Jan 1973. [13] D. Kim, T. Krishnamohan, L. Smith, H. S. P. Wong, and K. C. Saraswat, “Band to Band Tunneling Study in High Mobility Materials: III-V, Si, Ge and strained SiGe,” in Proceedings Device Research Conference, 2007, pp. 57–58. [14] J. Yipeng, W. Kangliang, W. Taihuan, D. Gang, and L. Xiaoyan, “Comparison of bandto-band tunneling models in Si and SiGe junctions,” Journal of Semiconductors, vol. 34, no. 9, pp. 092 002–1–092 002–5, Sept 2013. [15] “NSM Archive - Physical Properties of Semiconductors,” Jan. 2016. [Online]. Available: http://www.ioffe.ru/SVA/NSM/Semicond/Ge/bandstr.html BIBLIOGRAPHY 81 [16] P. Rastogi, T. Dutta, S. Kumar, A. Agarwal, and Y. S. Chauhan, “Quantum Confinement Effects in Extremely Thin Body Germanium n-MOSFETs,” IEEE Transactions on Electron Devices, vol. 62, no. 11, pp. 3575–3580, Nov 2015. [17] T. Dutta, S. Kumar, P. Rastogi, A. Agarwal, and Y. S. Chauhan, “Impact of Channel Thickness Variation on Bandstructure and Source-to-Drain Tunneling in Ultra-Thin Body III-V MOSFETs,” IEEE Journal of Electron Device Society, vol. 4, no. 2, pp. 66–71, Mar 2016. [18] T. Dutta, P. Kumar, P. Rastogi, A. Agarwal, and Y. S. Chauhan, “Atomistic study of band structure and transport in extremely thin channel InP MOSFETs,” physica status solidi (a), vol. 213, no. 4, pp. 898–904, 2016. [19] G. Lee, J. S. Jang, and W. Y. Choi, “Dual-Dielectric-Constant Spacer Hetero-GateDielectric Tunneling Field-Effect Transistors,” Semiconductor Science and Technology, vol. 28, no. 5, pp. 1–5, 2013. [20] E.-H. Toh, G. H. Wang, L. Chan, G. Samudra, and Y.-C. Yeo, “Device physics and guiding principles for the design of double-gate tunneling field effect transistor with silicongermanium source heterojunction,” Applied Physics Letters, vol. 91, no. 24, pp. 243 505– 1–243 505–3, 2007. [21] M. Zhu, S. S. Tan, E. H. Toh, and E. Quek, “Method of Fabricating a Silicon Tunneling Field Effect Transistor (TFET) with High Drive Current,” Apr 14 2011, US Patent 12 587 511. [22] “QuantumWise ATK 2014.2,” Jan. 2015. [23] Z. Jiang, J. Wang, H. H. Park, A. T. Pham, N. Xu, Y. Lu, S. Jin, W. Choi, M. A. Pourghaderi, J. Kim, and K. H. Lee, “Comprehensive Simulation Study of Direct Sourceto-Drain Tunneling in Ultra-Scaled Si, Ge, and III-V DG-FETs,” IEEE Transactions on Electron Devices, vol. 64, no. 3, pp. 945–952, Mar 2017. BIBLIOGRAPHY 82 [24] K. Sakata, B. Magyari-Kpe, S. Gupta, Y. Nishi, A. Blom, and P. Dek, “The effects of uniaxial and biaxial strain on the electronic structure of germanium,” Computational Materials Science, vol. 112, Part A, pp. 263 – 268, 2016. [25] A. N. Kholod, Saúl, Andrés, Fuhr, J. D., Borisenko, V. E., d’Avitaya, and F. Arnaud, “Electronic properties of germanium quantum films,” Phys. Rev. B, vol. 62, pp. 12 949– 12 954, Nov 2000. [26] E. O. Kane, “Theory of Tunneling,” Journal of Applied Physics, vol. 32, no. 1, pp. 83–91, 1961. [27] Bimberg, D., Blachnik et al., Physics of Group IV Elements and III-V Compounds / Physik Der Elemente Der IV. Gruppe und Der III-V Verbindungen, ser. Landolt-Bornstein: Numerical Data and Functional Relationshi. Springer, 1981. [28] C. Rivas, “Full bandstructure modeling of indirect phonon assisted transport in tunnel devices with silicon contacts,” Ph.D. dissertation, Dept. Elect. Engineering, Univ. Texas Dallas, USA, May 2004. [29] M. V. Fischetti, T. P. O’Regan, S. Narayanan, C. Sachs, S. Jin, J. Kim, and Y. Zhang, “Theoretical Study of Some Physical Aspects of Electronic Transport in nMOSFETs at the 10-nm Gate-Length,” IEEE Transactions on Electron Devices, vol. 54, no. 9, pp. 2116– 2136, Sept 2007. [30] K. H. Kao, A. S. Verhulst, W. G. Vandenberghe, B. Soree, G. Groeseneken, and K. D. Meyer, “Direct and Indirect Band-to-Band Tunneling in Germanium-Based TFETs,” IEEE Transactions on Electron Devices, vol. 59, no. 2, pp. 292–301, Feb 2012. [31] A. Rahman, M. S. Lundstrom, and A. W. Ghosh, “Generalized effective-mass approach for n-type metal-oxide-semiconductor field-effect transistors on arbitrarily oriented wafers,” Journal of Applied Physics, vol. 97, no. 5, pp. 053 702–1–053 702–12, 2005. [32] K. H. Kao, A. S. Verhulst, W. G. Vandenberghe, B. Soree, W. Magnus, D. Leonelli, G. Groeseneken, and K. D. Meyer, “Optimization of Gate-on-Source-Only Tunnel FETs BIBLIOGRAPHY 83 With Counter-Doped Pockets,” IEEE Transactions on Electron Devices, vol. 59, no. 8, pp. 2070–2077, Aug 2012. Chapter 5 Vertical and Lateral Tunneling in Ge Source TFET: Thickness Scaling 5.1 Introduction In this chapter, we present a comparative study of vertical and lateral tunneling in Ge source TFETs, as the source is scaled down from bulk to ultra-thin slabs of 4 nm thickness. It is well known that quantum confinement changes the band-structure and material parameters significantly. This has been reported in the literature for Ge and other III-V materials [1–3]. In bulk Ge, the difference between direct and indirect conduction band valleys (valley offset) is 0.14 eV [4], which reduces to 0.1 eV in ultra-thin Ge slabs [1]. The bandgap and effective masses for each valley are also known to change significantly. Additionally, at this finite thickness there will be a contribution of band-to-band tunneling (BTBT) current from both of the valleys on account of very low valley offset. Thus, we have theoretically calculated BTBT parameters for ultra-thin thickness using band structure results from atomistic simulations, while for bulk we have considered bulk parameters. In the investigated double gate TFET (DGTFET) structures, we have used an overlapped gate over Ge source to include vertical line tunneling along with lateral tunneling in the device. First, the impact of the gate overlap length scaling has been investigated at constant Ge source thickness of 8 nm. We have then analyzed the impact of scaling the Ge source thickness keeping the gate overlap length fixed. We find that Ge source thickness scaling till 8 nm has negligible effect on the ON current, but as the thickness is further 84 5.2 Device Structure and Simulation Models 85 scaled down to ultra-thin Ge films, line tunneling disappears in the device resulting in reduction of ION from 380 µA/µm at 25 nm (bulk) to 83.5 µA/µm at 4 nm. We also report TFETs with ultra-thin Ge source, but small gate-source overlap to confine the tunneling in Ge only. In this configuration, lateral tunneling drives the drain current, leading to better sub-threshold slope and ON current. Further, we highlight the roles of enhanced subsurface BTBT current and gate screening in these ultra-thin lateral TFETs. In the conventional TFETs, gate is usually aligned with the p+/n- junction. Tunneling through this junction from source to the channel is referred to as lateral point-to-point (P-P) tunneling. When there is gate-source overlap, the overlapped part of the source gets depleted. And, if a strong band bending is provided, then tunneling will take place from the bulk towards the gate/oxide interface. This vertical tunneling is referred as line tunneling. Line tunneling in TFETs has recently been explored to improve the ON current [5]. In this chapter we investigate the impact scaling the Ge source thickness from bulk to ultra-thin films for gate overlapped structure, accounting for both direct and indirect type of tunneling mechanisms as well as both line and P-P tunneling. The rest of the chapter is organized as follows: In section-5.2, we discuss the gate source overlapped structure along with the simulation models. The scaled Kane BTBT parameters for Ge source are discussed in section 5.3. This is followed by the discussion of the impact of gate overlap length (LOVL ) and thickness of Ge (TGe ) scaling in TFETs with gate-source overlap in section 5.4. The P-P lateral TFETs are discussed in section 5.5. Finally, we summarize our findings in section 5.6. 5.2 Device Structure and Simulation Models The TFET device structure under study is shown in Fig. 5.1. It has a gate-source overlap region (of length LOVL ), where vertical tunneling takes place. As the Ge source thickness (TGe ) reduces, the band structure changes due to quantum confinement effects, leading to increase in the direct and indirect bandgaps and the effective masses, and decrease in the conduction band valley-offsets. For TGe > 8 nm, the results are close to the bulk values, and hence we have con- 5.3 BTBT Model Parameters Calculations 86 Figure 5.1: Schematic of a double gate TFET with gate-source overlap, n-type drain doping of 1 × 1018 cm−3 , p-type source doping of 1 × 1020 cm−3 , and low n-type body doping of 1 × 1016 cm−3 . The gate dielectric is HfO2 with 2 nm thickness. The nominal gate-source overlap length (LOVL ) and the channel length are 8 nm and 40 nm respectively. The metal gate work function is set to 4.2 eV. sidered the bulk BTBT model parameters for thickness above 8 nm. For ultra-thin Ge source with TGe ≤ 8 nm, we have calculated the thickness dependent BTBT model parameters using the thickness dependent bandgap and effective mass data using density functional theory (DFT) calculations from our previous work [1]. The BTBT current simulations have been performed using the Synopsys Sentaurus TCAD suite, where we have included all the relevant physical effects. The non-local BTBT tunneling model [6] is used for the BTBT current, considering both direct and indirect tunneling in the device. The default values of BTBT model parameters for Ge in Sentaurus are available for indirect tunneling only, using bulk Ge properties. To overcome this problem at ultra thin Ge thickness, we have theoretically calculated the BTBT parameters for both direct and indirect tunneling for different Ge thickness. To include non-local electric field effects, hydrodynamic transport is considered and to account for the very high doping in source, bandgap narrowing and Fermi Dirac statistics are used. The geometrical quantum confinement effects in the BTBT are considered by explicitly calculating the BTBT model parameter after considering the thickness dependent geometrical confinement effects from atomistic simulations. To include the charge centroid shift due to the quantum effects, the MLDA model has been used. 5.3 BTBT Model Parameters Calculations According to the Kane dispersion relationship, the BTBT current can be defined as the integral of the generation rate over the entire device region [6] as already discussed in detail in section 5.4 Impact of LOVL and TGe Scaling in TFETs with gate-source overlap 87 Table 5.1: Theoretically calculated BTBT model parameters for direct and indirect tunneling for Ge (100) surface orientation with varying thickness. A and B are in units of cm−3 s−1 and MVcm−1 , respectively. Thickness Bulk 8 nm 6 nm 4 nm Direct BTBT Adir Bdir 20 1.47 × 10 6.04 1.46 × 1020 6.14 1.43 × 1020 6.6 1.35 × 1020 7.8 Indirect BTBT Aind Bind 15 1.67 × 10 6.55 1.509 × 1015 7.00 1.368 × 1015 7.62 1.163 × 1015 9.07 3.4. A and B are the Kane parameters, which for direct and indirect tunneling and are discussed in section 4.3.2 In Table 5.1, we have listed all the calculated BTBT model parameters for bulk and ultra-thin Ge. First, as a consistency check we note that for bulk Ge (100) the deviation between our theoretically calculated value for Bdir for direct tunneling (6.04 MV/cm) and the prediction from [8] (5.3 MV/cm) and the experimental calibration (5.7 MV/cm) [9] is small. It can be seen that in ultra-thin Ge slabs, the value of the parameter A decreases, while that of parameter B increases with thickness scaling. For BTBT parameters calculated using DFT bandstructure data, the value of Bdir at TGe = 8 nm increases from 5.7 MV/cm to 7.44 MV/cm at at TGe = 4 nm, which should degrade the current, because of the negative exponential dependence on B. The default values of Kane’s parameters used for Ge in Sentaurus TCAD are A = 9.1 × 1016 cm−3 s−1 and B = 4.9 MV/cm [10], which grossly overestimate the current at reduced thickness. 5.4 Impact of LOVL and TGe Scaling in TFETs with gate-source overlap First, we discuss the effect of the gate-source overlap length (LOVL ) on line tunneling at 8 nm source thickness. This is followed by a study of the thickness dependence of line tunneling for a constant LOVL of 8 nm. Then we investigate tunneling mechanisms in ultra-thin Ge source TFETs including the impact of change in bandstructure. 5.4 Impact of LOVL and TGe Scaling in TFETs with gate-source overlap 88 Figure 5.2: Transfer characteristics of DGTFET with different gate-source overlap lengths (LOVL ). The vertical line tunneling increases with LOVL , because of increased band bending provided by the overlapped part. Inset shows the ION and SSAVG variation with LOVL . ION increases linearly with LOVL . SSAVG remains almost constant with LOVL , but sharply increases at very low LOVL , because of increase in OFF current. SSAVG is calculated as the inverse of slope between VT (VGS at which IDS = 1 × 10−7 A/µm) and VOFF . 5.4.1 Effect of Overlap Length Scaling on Line Tunneling In this section, we have studied the effect of different gate overlap lengths (LOVL ) on line tunneling in TFETs as shown in Fig. 5.2. As LOVL increases, the amount of band bending increases, and as a result, the vertical tunneling also increases. The onset of vertical tunneling is shifted towards the left on the VGS axis with increase in LOVL . The ON current variation is almost linear with LOVL as shown in the inset of Fig. 5.2. The average subthreshold slope, SSAVG remains unaffected with decreasing LOVL except at 4 nm. This is because line tunneling sets in the device after the sub-threshold region and before this it’s lateral P-P tunneling which is playing in sub-threshold region. The adverse effect of increasing the LOVL is on capacitances, which increases with increasing LOVL . However, the increase in ON current is large as compared to the gate capacitances, thus, the switching speed and RF performance will improve. In the following sub-sections, we have considered a fixed LOVL of 8 nm in the devices. 5.4 Impact of LOVL and TGe Scaling in TFETs with gate-source overlap (a) 89 (b) (c) Figure 5.3: (a) Transfer characteristics of DGTFET with gate-source overlap, LOVL = 8 nm. Lateral P-P tunneling takes place initially at low VGS , and line tunneling sets in at higher VGS . Both, lateral and vertical tunneling will contribute to the total current. The inset shows ION and SSAVG for thickness scaling from bulk to ultra-thin Ge source. The ON current and SSAVG essentially remain constant as thickness is scaled down from bulk to 8 nm. For Ge thickness less than 8 nm, the ON current and SSAVG degrade, because of the disappearance of line tunneling (see Fig. 5.3a for more details).(b) Energy band diagrams are shown for different values of VGS at VDS = 0.7 V for 8 nm thickness along the lateral direction considered at 1 Å from the interface. Bands are pulled down with increase in VGS , which improves lateral P-P tunneling.(c) Energy band diagram along the vertical direction. As VGS increases, bands are pulled down and at around 0.5 V, vertical tunneling path is formed in the device, and this initiates line tunneling. 5.4.2 Bulk Ge source Line Tunnel FET (8 nm < TGe ≤ 25 nm) The total current in the structure considered in Fig. 5.1 is shown in Fig. 5.3a with the inset showing the variation of ION and SSAVG for different TGe values. The spatial view of energy band diagram along the vertical and horizontal directions for Ge source of thickness 8 nm (bulk) are shown in Fig. 5.3b and Fig. 5.3c, respectively. As mentioned earlier, the total current is a 5.4 Impact of LOVL and TGe Scaling in TFETs with gate-source overlap 90 Figure 5.4: Transfer characteristics for ultra-thin body Ge source having thickness less then 8 nm. The current degrades in the line tunneling configuration as thickness is scaled to 4 nm. Inset shows the energy band diagram along the vertical direction for ultra-thin Ge source at VGS = 1 V and VDS = 0.7 V. Gate voltage is not sufficient to provide enough band bending needed to initiate line tunneling in the device, and hence the total current is reduced. combination of vertical line tunneling and lateral P-P tunneling current, which in turn consist of both, direct and indirect tunneling. For bulk Ge source, at OFF state (VGS ≤ 0 V), there is no significant tunneling observed in the device, as the tunneling path associated with the lateral tunneling is large and also, there is no vertical tunneling path formed in the device, as depicted in Fig. 5.3b and Fig. 5.3c. Hence, the Shockley-Read-Hall (SRH) and BTBT current will constitute the OFF state current, and is shown in Fig. 5.3a. As VGS increases beyond 0 V, band overlapping starts which results in the formation of lateral tunneling path in the device. This tunneling path reduces with the increase in VGS , see Fig. 5.3b. Thus, initially it is lateral P-P tunneling, which contributes to the current in the device. However, as VGS reaches 0.5 V, there is sufficient band bending to initiate tunneling in the vertical direction, shown in Fig. 5.3c. This addition of line tunneling to the lateral P-P tunneling boosts the current after VGS = 0.5 V and hence the total current increases. As TGe is scaled down from 25 nm to 8 nm, the variation in ON current is seen to be very less, except for a slight decrease in the ON current due to the change in electrostatics in the structure, as shown in inset of Fig. 5.3a. The average sub-threshold slope SSAVG essentially remains constant with thickness scaling upto 8 nm. 5.4 Impact of LOVL and TGe Scaling in TFETs with gate-source overlap 91 Figure 5.5: Transfer characteristics of DGTFET with lateral P-P Tunneling. There is a cross over point in the transfer characteristics, which is due to the effect of gate coupling and gate screening. Inset shows the structure of lateral P-P TFETs. Source is extended inside the gate by 2 nm to limit the tunneling in Ge only. The hetero dielectric gate is composed of SiO2 (LSiO2 = 35 nm) and HfO2 (LHfO2 = 5 nm). ION and SSAVG variation with thickness is shown in the inset. ION degrades at reduced thickness of Ge source because of gate screening. 5.4.3 Ultra-Thin Ge Source Line Tunnel FET (TGe ≤ 8 nm) As the thickness is scaled down below 8 nm, the ON current is degraded by an order of magnitude and is shown in Fig. 5.4. This current degradation occurs due to the combined effect of disappearance of line tunneling and change in the electrostatic in the structure. For TGe ≤ 8 nm, the material confinement as well as external quantum effects in the device result in a very large increase in the band-gap for both direct and indirect valleys. The applied voltage at the gate is not sufficient to cause the voltage drop that is needed to initiate line tunneling in the device, and thus the current is degraded. At 4 nm thickness with LOVL = 8 nm, the gate has a better control over the source. But, due to the overlap of the gate with the heavily doped p-type source, depletion takes place in the overlapped source region, which increases the barrier width and thus causes the lateral P-P current to degrade, as shown in Fig. 5.4. Even a slight variation in VGS will result in heavy depletion in the source at 4 nm thickness, because of strong gate coupling and this leads to a degradation in the current. One way to circumvent this issue is to use low-κ gate dielectric to reduce the gate coupling, but it will also result in lesser ION . For 6 nm thickness, the positive effect of increased gate coupling and the adverse effect of increase in band-gap compensate each other. 5.5 Ultra-Thin Ge Source Lateral (P-P) TFET 92 This results in almost the same current for both thickness (6 nm and 8 nm) till 0.4 V. However, beyond 0.4 V, the current degrades because of gate screening. 5.5 Ultra-Thin Ge Source Lateral (P-P) TFET To overcome the problem of absence of line tunneling in ultra-thin Ge TFETs, we have studied a different lateral P-P tunneling structure, as shown in the inset of Fig. 5.5. In this device, we have used a hetero dielectric gate. Ge source is extended inside the gate by 2 nm to limit the tunneling to the ultra-thin Ge source only as Ge BTBT parameters are better than Si. We have considered the same device dimensions as the TFET device in the previous section. We have find that there is a cross-over behavior in the transfer characteristics with respect to thickness at VGS = 0.6 V, which is marked in Fig. 5.5. It can be explained as follows: For 0 V < VGS ≤ 0.6 V, reduced thickness results in stronger coupling of the gate to the body and sub-surface tunneling is enhanced. Accordingly, the OFF state current increases with reduced thickness: IOFF, 4nm > IOFF, 6nm > IOFF, 8nm as shown in Fig. 5.5. This sub-surface BTBT increases the sub-threshold current and there is substantial contribution to the total current for ultra-thin Ge source (TGe ≤ 8 nm). As the gate voltage is increased beyond 0.6 V, due to the formation of a strong tunneling region, the surface charge screens the gate voltage. This strong gate screening degrades the current for reduced thickness, and a cross-over happens. In the inset of Fig. 5.5, we have shown the variation of ON current and SS with thickness. The ON current variation with bulk thickness is only nominal, which is due to the slight change in the electrostatics of the structure (e.g. ION is found to be 241 µA/µm and 249 µA/µm in 10 nm and 20 nm devices, respectively). Comparatively, the ON current in TFETs with gate-source overlap (significant line tunneling) is around 52.2% higher than the ON current in lateral P-P TFETs at a thickness of 20 nm. However, as we approach ultra thin Ge thickness (TGe ≤ 8 nm), the reduction of line tunneling current in the gate-source overlap TFET structure degrades the total current, and ION is in fact only a fraction of the corresponding lateral P-P TFET current at the same thickness (e.g. the fraction being 54.5% and 56.2% at 4 nm and 6 nm thickness, respectively). 5.6 Summary 5.6 93 Summary We have shown that line tunneling disappears in the gate-source overlap TFETs as the Ge source thickness is scaled down to ultra-thin films (TGe ≤ 8 nm) and the net current is degraded. Thus, gate overlapped structure with ultra-thin Ge source is not suitable for future low power applications. We have shown that TFETs with strong lateral P-P tunneling results in better performance at reduced thickness as compared to the gate-source overlap TFETs that offer significant line tunneling only in the bulk case. The calculated ION varies from 312 µA/µm to 149.7 µA/µm for lateral P-P TFETs, while for line tunnel FETs with significant gate-source overlap, the ON current varies from 360 µA/µm to 83.5 µA/µm, as the source is scaled from 8 nm to 4 nm, respectively. Bibliography [1] P. Rastogi, T. Dutta, S. Kumar, A. Agarwal, and Y. S. Chauhan, “Quantum Confinement Effects in Extremely Thin Body Germanium n-MOSFETs,” IEEE Transactions on Electron Devices, vol. 62, no. 11, pp. 3575–3580, 2015. [2] T. Dutta, S. Kumar, P. Rastogi, A. Agarwal, and Y. S. Chauhan, “Impact of Channel Thickness Variation on Bandstructure and Source-to-Drain Tunneling in Ultra-Thin Body III-V MOSFETs,” IEEE Journal Electron Device Society, vol. 4, no. 2, pp. 66–71, 2016. [3] T. Dutta, P. Kumar, P. Rastogi, A. Agarwal, and Y. S. Chauhan, “Atomistic study of band structure and transport in extremely thin channel InP MOSFETs,” Physics Status Solidi (A), vol. 213, no. 4, p. 898904, 2016. [4] “NSM Archive-Physical Properties of Semiconductor.” [Online]. Available: http: //www.ioffe.ru/SVA/NSM/Semicond/Ge/bandstr.html [5] A. M. Walke, A. S. Verhulst, A. Vandooren, D. Verreck, E. Simoen, V. R. Rao, G. Groeseneken, N. Collaert, and A. V. Y. Thean, “Part I: Impact of Field-Induced Quantum Confinement on the Subthreshold Swing Behavior of Line TFETs,” IEEE Transactions on Electron Devices, vol. 60, no. 12, pp. 4057–4064, Dec 2013. [6] E. O. Kane, “Theory of Tunneling,” Journal of Applied Physics, vol. 32, no. 1, pp. 83–91, 1961. 94 BIBLIOGRAPHY 95 [7] K. H. Kao, A. S. Verhulst, W. G. Vandenberghe, B. Sore, G. Groeseneken, and K. D. Meyer, “Direct and Indirect Band-to-Band Tunneling in Germanium-Based TFETs,” IEEE Transaction on Electron Devices, vol. 59, no. 2, pp. 292–301, 2012. [8] D. Kim, T. Krishnamohan, L. Smith, H. S. P. Wong, and K. C. Saraswat, “Double-Gate Strained-Ge Heterostructure Tunneling FET (TFET) With record high drive currents and 60mV/dec subthreshold slope,” in Device Research Conference, 2007, pp. 57–58. [9] M. S. Tyagi, “Determination of Effective Mass and the Pair Production Energy for Electrons in Germanium from Zener Diode Characteristics,” Japanese Journal of Applied Physics, vol. 12, no. 1, pp. 106–108, 1973. [10] “Sentaurus Device, Synopsys, Version D-2010.03,” 2016. [Online]. Available: https://www.synopsys.com/silicon/tcad/device-simulation/sentaurus-device.html Chapter 6 Modeling of Current and Capacitances including Mobile Channel Charge and Ambipolar Behaviour 6.1 Introduction In this chapter, we have presented a surface potential based analytical model of charge, current and capacitances in DGTFET accounting for the mobile charges in the channel. To facilitate the use of TFETs in circuit design, an accurate analytical/compact model for TFET is needed, which is also computationally efficient. While there have been significant efforts to develop an analytical/compact model for TFETs, there is still considerable scope for improvement. A good TFET model should capture the device electrostatics accurately, and predict the ambipolar drain current along with the terminal charges and capacitances accurately in both the depletion as well as inversion regime, while being computationally efficient for circuit simulations. Earlier models based on 1-D Poisson equation do not accurately capture the electric field distribution in short channel devices [1]. Pseudo 2-D method based TFET models capture the electrostatics correctly, but some of them are valid only for depletion conditions [2–7], while some do not model the ambipolar current [2–4, 8] and some do not model the terminal charges and capacitances [4–9]. Additionally, there are some empirical drain current and capacitance model for TFETs as well [10–12], however these do not account for the mobile charges in the channel. 96 6.2 Device Structure and Simulation Models V V 97 G V 2 S o 3 1 2 nm o Drain Source (10 x y cm -3 R3 R2 R1 R1 p+(Si) 20 D Intrinsic Si n- (10 ) 16 t n+(Si) -3 16 -3 cm ) 10 cm ( 10 20 cm Si -3 ) L = 50 nm 2 x o x R1 x 1 R2 2 x R3 3 t Si = 10 nm Figure 6.1: Schematic of the DGTFET used in the study, where R1 is the source depletion region, R2 is the channel region, and R3 is the drain depletion region. VG , VS , and VD are the applied gate, source, and drain voltage, respectively and Φi represents the potential at the various marked positions. The doping in various regions and the channel length are mentioned in the figure. The workfunction of the metal at the gate is Φm = 4.2 eV. In this chapter, we present a surface potential based model for DG-TEFT, which uses the pseudo 2-D method to accurately model the surface potential in both the depletion as well as the inversion regime, which in turn is used to model the drain current along with the terminal charges and capacitances. Influence of mobile charge carrier on the device electrostatics is important to account for the precise modeling of the device characteristics. The mobile charges in the channel bring about i) narrowing in the tunneling barrier width [9] and ii) saturation in the monotonically increasing mid surface potential. While there are some earlier works which include one of these effect [8, 9, 13], we show that including both these effects are essential for modeling the drain current and the terminal charges and capacitances accurately. The chapter is organized as follows: In Sec. 6.2, we discuss the device structure and simulation details, followed by a detailed description of the model in Sec. 6.3, which in turn is followed by Sec. 6.4 on model validation and a summary of the work in Sec. 6.5. 6.2 Device Structure and Simulation Models 6.2 98 Device Structure and Simulation Models The TFET device structure under study is shown in Fig. 6.1. For simplicity, all the junctions in the device i.e. source/channel and channel/drain are assumed to be abrupt. The device is simulated using Sentaurus TCAD to obtain the simulation data for model verification [14]. In order to include the field and doping dependence mobility degradation, high field saturation and Masetti model is used along with the Lombardi Model. Shockley-Read-Hall (SRH), Boltzmann statistics, and Quasi static approximation is also activated. To calculate BTBT current, non local tunnelling path model is included [15]. 6.3 Model Description To calculate the band to band tunneling current in the device, we need to evaluate the electric field distribution throughout the device. To this end, we first calculate the surface potential in the device. The surface potential is determined primarily by the electrostatics of the system which is treated in 6.3.1, along with the impact of mobile charge carriers which is discussed in 6.3.2. The device is divided into three regions: R1, R2, and R3 as shown in Fig. 6.1. Pseudo 2-D method is adopted to convert the 2-D Poisson equation into an effective 1-D equation, which is further used to obtain the surface potential profile along the device. 6.3.1 Surface potential The 2-D Poisson equation without considering the effect of the mobile charges in the channel can be written as: ∂ 2 Φ(x, y) ∂ 2 Φ(x, y) −qNR + = , 2 2 ∂x ∂y Si (6.1) where Φ(x, y) is the electrostatic potential, q is the charge, Si is the electrical permittivity of silicon, and NR represents the doping in different regions of the device. The doping NR = NS , Nch , and ND for region R1 , R2 , and R3 , respectively. To proceed further and to convert (6.1) into 6.3 Model Description 99 an effective 1-D equation, we consider a parabolic approximation for the electrostatic potential along the thickness of the device [2], Φ(x, y) = a0 (x) + a1 (x)y + a2 (x)y 2 , (6.2) where a0 (x), a1 (x), and a2 (x) are the unknown coefficients which depend on x. The value of these unknown coefficients can be obtained using the boundary conditions for the electrostatic potential and electric field in (6.2). The effective gate voltage (Φg ) is given by Φg = VG − Φm + χ + Eg , 2q (6.3) where VG is the applied gate voltage, Φm is the metal gate workfunction, χ is the electron affinity, and Eg is the energy bandgap. The symmetric double gate structure of the device implies that the surface potential at the front as well as the back gate are equal, Φ(x, 0) = Φ(x, tSi ) ≡ Φs (x). Furthermore, electric field at the front surface (E(x, 0)) and back surface (E(x, tSi )) can be expressed as follows [2]: Cox [Φg − Φs (x)] , Si Cox [Φs (x) − Φg ] E(x, tSi ) = . Si E(x, 0) = (6.4) In (6.4), Cox is the gate oxide capacitance per unit area and is equal to ox /tox for region R2 and ox /t for region R1 and R3 with t = πtox /2. Note that t is the effective oxide thickness in the R3 region, which accounts for the effect of fringing fields of the gate [16]. Now applying the boundary conditions from (6.4) into (6.2), we obtain the unknown coefficients of (6.2) as a0 (x) = Φs (x), a2 (x) = − a1 (x) = Cox (Φg − Φs (x))/Si tSi . tsi (6.5) Now using a0 (x), a1 (x), and a2 (x) from (6.5) into (6.2) and substituting Φ(x, y) in (6.1), we 6.3 Model Description 100 obtain an effective 1-D Poisson equation, ∂ 2 Φs,i (x) (0) − ki2 Φs,i (x) = −ki2 Φd,i , ∂y 2 (6.6) where i = 1, 2, and 3 correspond to regions R1, R2, and R3, respectively, and r ki = (0) qNR tSi 2Cox (0) , Φd,i = Φg + . Si tsi 2Cox (6.7) (0) Here Φd,i in region R2 i.e. Φd,2 is called the center or the mid potential under depletion condition and it is a linear function of applied gate voltage and ki is the inverse decay length in the depletion condition. Note that till now we have not included the effect of mobile charge carriers (0) in the channel region and this will lead to the modification of Φd,2 and k2 as discussed later in Sec. 6.3.2. The general solution of (6.6) can be expressed as: Φs,i (x) = bi eki (x−xi−1 ) + ci e−ki (x−xi−1 ) + Φd,i , (6.8) where xi−1 is position x0 , x1 , x2 , and x3 corresponding to i = 0, 1, 2, and 3 as defined in Fig. 6.1. Defining Φs,i (xi )=Φi , the unknown coefficients bi and ci of (6.8) can be evaluated using the following boundary conditions, Φ0 = −Φt ln NS ni = b1 + c1 + Φd,1 , Φ1 = b1 ek1 L1 + c1 e−k1 L1 + Φd,1 = b2 + c2 + Φd,2 , (6.9) Φ2 = b2 ek2 L2 + c2 e−k2 L2 + Φd,2 = b3 + c3 + Φd,3 , ND Φ3 = Vd + Φt ln = b3 ek3 L3 + c3 e−k3 L3 + Φd,3 , ni By solving the linear equations specified in Eq. 6.9. The coefficient b1 , c1 , b2 , c2 , b3 and c3 6.3 Model Description 101 are obtained, and are given by: b1 = 1 φ1 − φd1 − e−k1 L1 (φ0 − φd1 ) , 2sinh(k1 L1 ) (6.10) c1 = k1 L1 1 e (φ0 − φd1 ) − φ1 + φd1 , 2sinh(k1 L1 ) (6.11) b2 = k2 L2 1 (φ1 − φd2 ) − φ2 + φd2 , e 2sinh(k2 L2 ) (6.12) c2 = 1 φ2 − φd2 − e−k2 L2 (φ1 − φd2 ) , 2sinh(k2 L2 ) (6.13) b3 = 1 φ3 − φd3 − e−k3 L3 (φ2 − φd3 ) , 2sinh(k3 L3 ) (6.14) c3 = k3 L3 1 e (φ2 − φd3 ) − φ3 + φd3 . 2sinh(k3 L3 ) (6.15) Here, Φ0 and Φ3 are the surface potential at the source and drain end, respectively, and Φ1 and Φ2 are the surface potential at the boundary between R1/R2 and R2/R3, respectively as marked in Fig. 6.1, ni denotes the intrinsic charge carrier concentration, while the thermal voltage is given by Φt = kT /q. In (6.9), L1 and L3 are the depletion region length in the source and drain region, respectively, and L2 is the channel length. Within the depletion approximation, L1 and L3 are given by s L1 = 2Si (Φ1 − Φ0 ) , L3 = qNS s 2Si (Φ3 − Φ2 ) . qND (6.16) Note unlike some earlier works [2], (6.16) includes the physically relevant potential drop across the junction in calculation of depletion length. Finally we can evaluate Φ1 and Φ2 at the boundary of R1/R2 and R2/R3 using continuity of the potential following [17] and expressed as: q Φ1 = Φd,2 + NS α2 − 1.7NS (Φd,2 − Φ0 )α2 + NS2 α22 , (6.17) q 1.7ND (Φ3 − Φd,2 )α2 + ND2 α22 , (6.18) and Φ2 = ND α2 − Φd,2 − 6.3 Model Description 102 1.4 V = 0.8 V DS Symbols: Simulation Line: Model V = 0.6 V DS (V) 1.2 V = 0.4 V DS d,2 1.0 V = 0.2 V DS 0.8 0.6 0.4 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 V (V) GS Figure 6.2: Plot of the channel mid potential as a function of VGS for different values of VDS . The linear portion corresponds to the depletion in the channel, while the saturated region indicates inversion in the channel. The model results are consistent with the TCAD simulation data. where we have defined α2 = q k22 Si tanh2 (L2 k2 ) . (6.19) This gives us a closed form analytical expression for the surface potential and the full spatial potential profile along the channel of the device can be obtained. 6.3.2 Effect of mobile charges on Φs and k2 : dual modulation effect In this subsection, we include the effect of mobile charges on the potential profile in the device. The mobile charges in the channel bring about i) narrowing in the tunneling barrier width and ii) saturation in the monotonically increasing mid surface potential and this is typically referred to as the dual modulation effect. The narrowing of the tunnel barrier influences the band to band tunneling (BTBT) current [8, 13], while the saturation of the mid or center surface potential is essential for accurate prediction of terminal charges and therefore capacitances [9]. The saturation in the mid surface channel potential is modeled via an empirical equation and the narrowing of tunneling barrier is included as a correction in the inverse decay length calculated using a variational approach. 6.3 Model Description 103 Incorporating the saturation effect in the mid-surface potential (Φd,2 ) [9] yields, Φd,2 q (0) (0) 2 2 = 0.5 Φd,2 + Φinv − (Φd,2 − Φinv ) + δ , (6.20) (0) where Φd,2 is the mid surface potential without the mobile charge carrier effect and (0) (0) Φinv = VD + Ψ + α(Φd,2 − VD − Ψ) + β(Φd,2 − VD − Ψ)2 . (6.21) In (6.21), α and β are the fitting parameters and Ψ is given by Nch Ntran Ψ = Φt ln , n2i (6.22) where Ntran is the doping corresponding to the transition from linear variation to saturation in the mid surface potential and its extracted value is 1014 cm−3 . In Fig. 6.2, we have shown the modeled Φd,2 in (6.20) against the TCAD simulation data. The linear portion in Fig. 6.2 corresponds to the depletion in the channel, while the saturated region shows the screening effect of the mobile charge carriers in the channel. To include the mobile charge effect in the inverse decay length, a variational approach for the effective action is adopted to solve the 2-D non-linear Poisson equation. Following Ref. [13], we have Z Z S(Φ) = 2 ∂Φ ∂Φ 2 dxdy + − ρΦ , 2 ∂x ∂y (6.23) where Φ(x, y) is the electrostatic potential and ρ(x, y) is the charge density. On minimizing S(Φ) and neglecting the y dependences, i.e., dS(Φ)/dk2 = 0 we obtain the modified (see Eq. (6.7)) inverse decay length to be s k2 = 2Cox αNinv − , Si tSi Si tSi (Φ0 − Φd,2 ) (6.24) where Ninv =2Cox (Φg -Φd,2 ) is the inversion charge in the channel and α is a fitting parameter whose value lies between 1-2. In (6.24), the second term in r.h.s. accounts for the mobile 6.3 Model Description 104 charge carrier effect and is derived for a long channel length device to decouple the effect of drain voltage [17]. Including the drain voltage dependence, the expression is further modified as follows: s k2 = 2Cox αNinv (1 + Vdseff ) − , Si tSi Si tSi (Φ0 − Φd,2 ) (6.25) where (1 + Vdseff ) is added to include the drain voltage effect. In (6.25), Vdseff implies Vds in the linear region and Vdsat in the saturation region and in order to ensure smooth transition from linear to saturation region it is modeled as [18]: 1 Vdseff = Vdsat − 2 q 2 (Vdsat − Vds − Υ) + (Vdsat − Vds − Υ) + 4ΓVdsat (6.26) where Υ and Γ are fitting parameters [18]. The calculation of Vdsat is similar to that in conventional MOSFET, and it is given by Vdsat q 2 = Φg + γ 2 /4 − γ/2 − Φ , where Φ and γ are kT Φ= log q N2 Nm n2i √ , γ= 2Si qN2 . Cox (6.27) (6.28) In (6.28), Nm = 1.8 × 1018 cm−3 . 6.3.3 Electric Field The electric field in the source and drain depletion region can be approximated by lateral electric field by ignoring the fringing field effect from the gate. However inside the channel both the lateral and the vertical component of the field have to be considered. The vertical component of electric field (Ex ) can be obtained by differentiating (6.2) as Ex (x, y) = a1 (x) + 2a2 (x)y , (6.29) 6.3 Model Description 105 and the lateral component of electric field (Ey ) can be evaluated by differentiating (6.8) and is given by Ey (x, y) = ki bi eki (x−xi−1 ) − ki ci e−ki (x−xi−1 ) . (6.30) Finally, magnitude of the total electric field (Exy ) is computed as q Exy = Ex2 (x, y) + Ey2 (x, y) . (6.31) The electric field predicted by the model in (6.31) deviates from the simulation data [2]. In order to compensate the mismatch, a smoothing function in the electric field is used as follows [19]: where E = Exy (1 − f ) + ζExy f , (6.32) 22 Φg − Φt ln 10 Nch . f = 0.5 1 + tanh 4Φt (6.33) In (6.33), ζ is a fitting parameter and f is a empirical function and its value varies between 0 to 1 [19]. 6.3.4 BTBT Current The BTBT current is defined as the integral of the generation rate over the entire device region. In presence of a uniform electric field, BTBT current from source to channel (Itun,s→c ) in Ampere/µm can be expressed as follows [15]: Z tSi Z Wtmin,s Itun,s→c = q 0 Wtmax,s P E B A exp − dydx , E0 E (6.34) where E0 = 1 V/cm, P = 2 and 2.5 for the direct and indirect tunneling, respectively, A and B are the parameters for the Kane model with the unit of cm−3 s−1 and M V cm−1 , respectively and E is the electric field at a given location across the tunneling junction. The values of A and B that we have considered in our work are 4 × 1014 cm−3 s−1 and 1.9 × 107 M V cm−1 , respectively. 6.3 Model Description 106 1.5 0.5 (a) Wtmin,s Eg 0.0 Wtmax,s -0.5 -1.0 -1.5 Source 1.0 Energy (eV) Energy (eV) 1.0 Channel (b) 0.5 Wtmax,d 0.0 Wtmin,d -0.5 -1.0 -1.5 -2.0 -2.5 0.00 0.02 0.04 0.04 0.06 Distance ( m) Drain Channel -2.0 0.06 Distance ( 0.08 m) Figure 6.3: Energy band profile in DGTFET showing minimum and maximum tunneling path at; a) source /body junction and b) drain/ body junction. In order to simplify the calculation, it is assumed that integration along the vertical direction is constant and the tunneling is taking place laterally along the surface. This assumption gives satisfactory result and is widely used in modeling of the current in TFETs [2, 9, 20]. Z Wtmin,s Itun,s→c = qAtSi Wtmax,s E E0 P B exp − dx , E (6.35) The range of energies over which tunneling takes place can be calculated through the evaluation of the minimum and the maximum tunneling path in the source and drain depletion region. The length over which the channel potential rises by Eg /q w.r.t. the source potential is the minimum tunneling path at source end (Wtmin,s ) (see Fig. 6.3a), while, the maximum tunneling path at source end (Wtmax,s ) is the width over which the channel potential falls by Eg /q w.r.t. the mid potential (see Fig. 6.3b). Using the above definition of minimum and maximum tunneling length, we can evaluate the tunneling lengths at source side to be 1 Eg Wtmin,s = ln(c2 ) − ln Φ1 + − Φd,2 , 2k2 q 1 Eg Wtmax,s = ln(c2 ) + ln . 2k2 q (6.36) Since the tunneling probability as specified by WKB (Wentzmer Krammer Brillouin) approximation decays exponentially with increase in tunneling path, it is expected that the generation 6.3 Model Description 107 rate in the Wtmin region is the maximum contribution to the BTBT tunneling current [21]. Following a similar approach, the BTBT tunneling current and tunneling distance at the drain side can be expressed as, Z Wtmin,d Itun,d→c = qAtSi Wtmax,d P E B A exp − dx , E0 E 1 Eg Wtmin,d = L2 − ln(c2 ) − ln Φ2 − − Φd,2 , 2k2 q 1 Eg Wtmax,d = L2 − ln(c2 ) − ln . 2k2 q (6.37) (6.38) Here, Wtmin,d is the minimum tunneling path along the lateral direction at the drain junction and it is the distance over which surface potential in channel falls by Eg /q w.r.t. drain potential (see Fig. 6.3). In (6.38), Wtmax,d is the maximum tunneling length at the drain junction and is the distance over which surface potential in the channel increases by a amount of Eg /q w.r.t. channel potential Φd,2 . The leakage current is included in the total current via an empirical equation as a function of effective gate voltage [19]: Ileakage Φg = I0 exp − , 7Φt (6.39) where I0 is the pre-factor used to capture the leakage current. Now, using tunneling current from source to channel and drain to channel, the total current is simply given by IDS = Itun,s→c + Itun,c→d + Ileakage . 6.3.5 (6.40) Terminal Charge and Capacitance To model the capacitance in the device, it is necessary to express the terminal charges variation w.r.t to terminal voltages. Unlike the conventional MOSFET, in homo-junction TFETs only the drain is connected to the inversion charge [22]. Therefore, entire channel charge is assumed to be associated with drain terminal and the charge associated with source terminal mainly con- 6.3 Model Description 108 sists of depletion charge. The charge associated with the device terminals i.e. drain charge (QD ) and source charge (QS ) are modelled using 100/0 charge partitioning scheme. A crude approximation of the uniform charge distribution along the channel is used, which gives satisfactory result for TFETs [20, 23]. Accounting for the inversion and depletion charge contribution, the drain terminal charge is expressed as QD = QDinv + QDdep , (6.41) QDinv = −2Wg Cox (LG − Wtmin )(Φg − Φd,2 ) , (6.42) where is the inversion charge at the drain end and Wg is the device width and QDdep = Si (VDS + Vbi − Φd,2 )tSi Wg k2 , (6.43) denotes the depletion charge at the drain end. Since the charge at the source terminal (QS ) is only due to depletion, therefore, depletion charge calculation approximation is used to evaluate QS as follows [10]: QS = −qWg NS L1 tSi . (6.44) Now using charge conservation, the total gate charge (QG ) is simply given by QG = −(QD + QS ) . (6.45) The terminal capacitances can now be calculated by differentiating the terminal charges w.r.t. applied voltages and can be expressed as, CGG = ∂QG −∂QD , CDG = . ∂VGS ∂VGS (6.46) 6.4 Model Validation 6.4 109 Model Validation The proposed model is validated against TCAD simulation data for wide range of applied gate and source-drain voltages. As we will see below, the model is in reasonable agreement with the simulation results in most of the operating regimes. 6.4.1 Surface Potential, Electric field, and Current The surface potential (Φs ) based on the developed model, excluding the mobile charge carrier effect is shown in Fig. 6.4a along with the TCAD simulation data. The surface potential matches well with the simulation data at lower VGS and at VDS = 1.0 V, corresponding to the depletion conditions in the channel but it starts showing some deviation close to VGS =1.2 V due to the presence of mobile charge in channel, where the saturation is not captured by the model equation for the depletion condition. This deviation at higher VGS values highlights the necessity for accounting for the effect of mobile charges on the center potential of the channel. From Fig. 6.4b, it is evident that the surface potential result of the model matches well with the simulation data for a wide range of the applied drain to source voltage (VDS ), when the device is in the depletion regime at VGS = 0.2 V. To overcome the model deficiency in the strong inversion condition, effect of mobile charges is included in the surface potential and its verification with the simulation data for different applied bias conditions is shown in Fig. 6.4c. In Fig. 6.4c, the surface potential is shown at VGS = 1.0 V (corresponding to strong inversion regime in device), for different VDS values. To obtain close match of model with the simulation data in the entire length of the device, improvement in the inverse decay length is further required. It is noticeable that pinning of the surface potential at the tunnel junction i.e. source/channel interface is precisely captured by the model. The BTBT current is mainly driven by the maximum electric field [21], which is at the tunnel junction. Deviation of the model in the vicinity of tunnel junction as shown in Fig. 6.4c affect the BTBT at low VDS , which is a shortcoming of the model. Verification of the total electric field along the channel with smoothing function is shown in Fig. 6.4d at VDS = 1.0 V, for gate voltages corresponding to depletion to strong inversion in the device. 110 VGS= 1.2 Symbol: Simulation VDS= 1 V (a) V Line: Model GS= -1 Surface Potential (V) Surface Potential (V) 6.4 Model Validation 2.0 Symbol: Simulation 1.5 1.0 0.5 (b) -0.5 1.0 0.5 (c) Symbols: Simulation Line: Model 0.03 0.04 0.05 m 0.06 Lateral distance ( ) 0.07 Electric Field (MV/cm) Surface Potential (V) VGS= 1.0 V -0.5 0.02 DS= 0 V 0.03 0.04 0.05 0.06 Lateral distance ( VDS= 0, 0.2, 0.4, 0.6, 0.8, 1 and1.2 V 0.0 V 0.0 m) 2.0 1.5 GS= 0.2 V V = 1.2 V DS 0.02 Lateral distance ( V Line: Model 0.07 m) 4.5 Symbol: Simulation Line: Model 3.6 V = 1 V DS 2.7 V = 0,0.4, 0.8, 1.2 V GS 1.8 (d) 0.9 0.0 0.02 0.03 0.04 0.05 0.06 Lateral distance ( 0.07 m) Figure 6.4: Validation of the surface potential and electric field behavior of model against the TCAD simulation data along the lateral distance in x-axis. In panel(a), the effect of mobile charge carrier is not included, while it is included in panels (b), (c) and (d). In panel (a), model without mobile charge effect shows excellent matching with simulation data at lower VGS but start deviating with increase in mobile charge carrier at higher gate voltage at VGS =1.2V. The model in depletion condition at VGS =0.2V in panel (b) has excellent match with simulation data. The deviation in panel (a) at higher VGS highlights the necessity to include the mobile charge carriers in the model, which corrects the deviation at high VGS values as shown in (c). The panel (c) displays the surface potential profile along the channel for VGS =1.0V and VDS =0.0V to 1.2V in steps of 0.2V. The saturation in the mid potential is effectively captured by the model along with the pinning of the surface potential at the source end. (d) Validation of the total electric field predicted by the model along the channel with TCAD simulation data in the channel for VDS = 1.0V and different VGS values with smoothing function. The surface potential obtained from the model, is further used to model the drain current and its validation with the simulation data is shown in Fig. 6.5. Transfer characteristics as shown in Fig. 6.5a highlight the closeness of the proposed drain current model to the TCAD simulation data in the linear as well as in the saturation regime. The ambipolar current predicted by the 6.4 Model Validation 10 IDS(A/ 10 10 10 10 10 10 10 Line: Model V = 1.0 V DS 2.0x10 = 1.4 V V GS -9 -11 V =0.2 V DS -12 -13 m) m) 10 Symbol: Simulation 1.5x10 IDS(A/ 10 -9 -10 111 (a) -15 Symbol: Simulation (b) Line: Model = 1.3 V V GS 1.0x10 -14 -9 -9 = 1.2 V V GS -16 5.0x10 -17 -10 V = 1.1 V GS = 0.2, 0.4, 0.6, 0.8 and 1V V DS -18 0.0 -0.5 0.0 0.5 1.0 0.0 1.5 0.2 0.4 VGS(V) 0.6 0.8 1.0 1.2 1.4 1.6 1.8 VDS(V) Gate Charge (QG){fC} Drain Charge (QD) {fC} Figure 6.5: Drain current predicted by the model verified with TCAD simulation data; (a) transfer characteristics at varying gate voltage for different VDS . (b) Output current characteristics at different VGS values, perfect saturation in the drain current is achieved due to the pinning of the surface potential at the source side. 0.4 1.6 VDS= 0, 0.2, 0.4, 0.6, 0.8 and 1V 0.0 -0.4 1.2 (a) 0.8 -0.8 VDS= 0, 0.2, 0.4, 0.6, 0.8 and 1 V Symbols: Simulation Line: Model (b) 0.4 Symbol: Simulation Line: Model -1.2 0.0 0.0 0.4 0.8 1.2 VGS(V) 1.6 0.0 0.4 0.8 1.2 1.6 VGS(V) Figure 6.6: Terminal charge (QD , and QG ) predicted by the model against simulation data; (a) Drain terminal charge for different VDS w.r.t. applied gate to source voltage and (b) Gate terminal charge for different VDS w.r.t. applied gate to source voltage. model qualitively follows the TCAD simulation data as shown in Fig. 6.5a. Note that effect of the phonon scattering on the OFF current [24] is not included in the presented model as well as in the TCAD simulation. Output characteristics in Fig. 6.5b show that the model reproduces the simulation data reasonably well. 1.2 112 1.2 VDS= 0, 0.2, 0.4, 0.6, 0.8 and 1 V 0.8 VDS= 0, 0.2, 0.4, 0.6, 0.8 and 1V 0.8 (a) 0.4 0.0 Gate Capacitance (C ) {fF} GG Drain Capacitance (C ) {fF} DG 6.5 Summary (b) 0.4 0.0 0.4 0.8 VGS(V) 1.2 1.6 Symbols: Simulation Line: Model 0.0 0.4 0.8 1.2 1.6 VGS(V) Figure 6.7: Validation of drain capacitance (CDG ) and gate capacitance (CGG ) behavior of model against the TCAD simulation data. The panel (a) shows the CDG while panel (b) shows the CGG at different values of VDS w.r.t. applied gate to source voltage. Note that CGG and CDG looks similar, but their values are different as QG includes QS in addition to the QD . 6.4.2 Terminal Charges and Capacitances The gate charge (QG ) and the drain charge (QD ) predicted by our model w.r.t. the applied VGS at different VDS are shown in Fig. 6.6a and Fig. 6.6b, respectively along with the simulation data. The terminal drain capacitance (CDG ) and gate capacitance (CGG ) w.r.t the applied VGS at VDS is shown in Fig. 6.7a and Fig. 6.7b and show a reasonable match with the simulation results. In both the figures, the model shows reasonable consistency with the simulation data. In general proposed model captures the simulation data trend satisfactorily, albeit with small deviation at higher VGS values. This small deviation may be due to effects such as fringing field capacitance, gate overlap capacitance etc. which are not included in our present work. Note that CGG and CDG looks similar, but their values are different as QG includes QS in addition to the QD . 6.5 Summary In this chapter, we have developed a surface potential based analytical model of the drain current, terminal charges, and capacitances for a DG-TFET. The surface potential model includes the effect of mobile charges and is valid over a wide range of applied bias conditions including 6.5 Summary 113 both depletion as well as the saturation regime. The surface potential model is in turn used to model the ambipolar current in the device. Furthermore, we also calculate the terminal charge and terminal capacitance. Our model shows a good agreement with the TCAD based simulation data for the electric field, drain current, terminal charges, and various terminal capacitances in a DG-TFET over a wide operating range. This highlights the fact that proposed model captures all the relevant aspects of the device physics reasonably well, and it can be very useful as a compact model for computationally efficient circuit simulations of a DG-TFET. The presented model methodology can also be adopted to model the III-V heterojunction TFETs by incorporating the effect of the heterojunction. Bibliography [1] V. Dobrovolsky and F. Sizov, “Analytical Model of the Thin-Film Silicon-On-Insulator Tunneling Field Effect Transistor,” Journal of Applied Physics, vol. 110, pp. 114 513–1– 114 513–7, 2011. [2] M. G. Bardon, H. P. Neves, R. Puers, and C. V. Hoof, “Pseudo-Two-Dimensional Model for Double-Gate Tunnel FETs Considering the Junctions Depletion Regions,” IEEE Transactions on Electron Devices, vol. 57, no. 4, pp. 827–834, Apr 2010. [3] S. Kumar, E. Goel, K. Singh, B. Singh, M. Kumar, and S. Jit, “A Compact 2-D Analytical Model for Electrical Characteristics of Double-Gate Tunnel Field-Effect Transistors With a SiO2 /High-k Stacked Gate-Oxide Structure,” IEEE Transactions on Electron Devices, vol. 63, no. 8, pp. 3291–3299, Aug 2016. [4] A. S. Verhulst, B. Sore, D. Leonelli, W. G. Vandenberghe, and G. Groeseneken, “Modeling the single-gate, double-gate, and gate-all-around tunnel field-effect transistor,” Journal of Applied Physics, vol. 107, no. 2, p. 024518, 2010. [5] V. Prabhat and A. K. Dutta, “Analytical Surface Potential and Drain Current Models of Dual-Metal-Gate Double-Gate Tunnel-FETs,” IEEE Transactions on Electron Devices, vol. 63, no. 5, pp. 2190–2196, May 2016. [6] M. J. Lee and W. Y. Choi, “Analytical Model of Single-Gate Silicon-on-Insulator (SOI) Tunneling Field-Effect Transistors (TFETs) ,” Solid-State Electronics, vol. 63, no. 1, pp. 110 – 114, 2011. 114 BIBLIOGRAPHY 115 [7] N. Cui, L. Liu, Q. Xie, Z. Tan, R. Liang, J. Wang, and J. Xu, “A Two-Dimensional Analytical Model for Tunnel Field Effect Transistor and Its Applications,” Japanese Journal of Applied Physics, vol. 52, no. 4R, p. 044303, 2013. [8] M. Gholizadeh and S. E. Hosseini, “A 2-D Analytical Model for Double-Gate Tunnel FETs,” IEEE Transactions on Electron Devices, vol. 61, no. 5, pp. 1494–1500, May 2014. [9] C. Wu, R. Huang, Q. Huang, C. Wang, J. Wang, and Y. Wang, “An Analytical Surface Potential Model Accounting for the Dual-Modulation Effects in Tunnel FETs,” IEEE Transactions on Electron Devices, vol. 61, no. 8, pp. 2690–2696, Aug 2014. [10] J. Wang, C. Wu, Q. Huang, C. Wang, and R. Huang, “A Closed-Form Capacitance Model for Tunnel FETs with Explicit Surface Potential Solutions,” Journal of Applied Physics, vol. 116, no. 9, pp. 094 501–1–094 501–8, 2014. [11] H. Lu, D. Esseni, and A. Seabaugh, “Universal Analytical Model for tunnel FET Circuit Simulation,” Solid-State Electronics, vol. 108, pp. 110–117, 2015. [12] S. Mookerjea, R. Krishnan, S. Datta, and V. Narayanan, “Effective Capacitance and Drive Current for Tunnel FET (TFET) CV/I Estimation,” IEEE Transactons on Electron Devices, vol. 56, no. 9, pp. 2092–2098, Sept 2009. [13] C. Shen, S. L. Ong, C. H. Heng, G. Samudra, and Y. C. Yeo, “A Variational Approach to the Two-Dimensional Nonlinear Poisson’s Equation for the Modeling of Tunneling Transistors,” IEEE Electron Device Letters, vol. 29, no. 11, pp. 1252–1255, Nov 2008. [14] “Sentaurus Device, Synopsys, Version D-2010.03,” 2010. [Online]. Available: https://www.synopsys.com/silicon/tcad/device-simulation/sentaurus-device.html [15] E. O. Kane, “Theory of Tunneling,” Journal of Applied Physics, vol. 32, no. 1, pp. 83–91, 1961. BIBLIOGRAPHY 116 [16] S. C. Lin and J. B. Kuo, “Modeling the Fringing Electric Field Effect on the Threshold Voltage of FD SOI nMOS Devices with the LDD/Sidewall Oxide Spacer Structure,” IEEE Transactions on Electron Devices, vol. 50, no. 12, pp. 2559–2564, Dec. 2003. [17] A. Pan and C. O. Chui, “A Quasi-Analytical Model for Double-Gate Tunneling FieldEffect Transistors,” IEEE Electron Device Letters, vol. 33, no. 10, pp. 1468–1470, Oct 2012. [18] “BSIM6 Technical Manual and Code,” Jul 2015. [Online]. Available: http: //www-device.eecs.berkeley.edu/bsim/ [19] A. Pal and A. K. Dutta, “Analytical Drain Current Modeling of Double-Gate Tunnel FieldEffect Transistors,” IEEE Transactions on Electron Devices, vol. 63, no. 8, pp. 3213–3221, Aug 2016. [20] L. Zhang, X. Lin, J. He, and M. Chan, “An Analytical Charge Model for Double-Gate Tunnel FETs,” IEEE Transactions on Electron Devices, vol. 59, no. 12, pp. 3217–3223, Dec 2012. [21] R. Vishnoi and M. J. Kumar, “Compact Analytical Model of Dual Material Gate Tunneling Field-Effect Transistor Using Interband Tunneling and Channel Transport,” IEEE Transactions on Electron Devices, vol. 61, no. 6, pp. 1936–1942, Jun 2014. [22] Y. Yang, X. Tong, L.-T. Yang, P.-F. Guo, L. Fan, and Y.-C. Yeo, “Tunneling FieldEffect Transistor: Capacitance Components and Modeling,” IEEE Electron Device Letters, vol. 31, no. 7, pp. 752–754, July 2010. [23] J. U. Mehta, W. A. Borders, H. Liu, R. Pandey, S. Datta, and L. Lunardi, “III -V Tunnel FET Model With Closed-Form Analytical Solution,” IEEE Transactions on Electron Devices, vol. 63, no. 5, pp. 2163–2168, May 2016. [24] M. G. Pala, C. Grillet, J. Cao, D. Logoteta, A. Cresti, and D. Esseni, “Impact of inelastic phonon scattering in the OFF state of Tunnel-field-effect transistors,” Journal of Computational Electronics, vol. 15, no. 4, pp. 1240–1247, 2016. Chapter 7 Conclusions Tunnel field effect transistor (TFET) is one of the promising devices that circumvent the subthreshold swing problem in modern low power transistors. TFETs have been widely researched recently for low power VLSI devices, because of it’s low OFF current and steeper subthreshold slope. The limitations of the TFET are its low ON-current as compared to MOSFET/FinFET requirements and stringent fabrications steps required for good performance. Reported methods to increase ON current in TFETs include structural modifications (hetero junction source), doping optimization (retrograde doping, pocket implanted source etc), gate metal work function engineering (double gate, dual metal gate), improving electrostatics (gate all around structures), and dielectric optimization (with one or multiple dielectrics). Strain at the source and drain end can also be used to increase the ON current. The lower band gap material at the source side in tunnel based devices favors better tunneling current. In order to tackle some of the issues associated with TFETs, we have investigated some optimization techniques in details. We have first discussed “Dual Metal Gate” technique, in which two metals are placed at the gate for Si and relaxed SiGe source. Dual metal at the gate provide additional flexibility to control the different sections of the transfer characteristics. Further, we have observed that the performance improves by manoeuvring the dielectrics at the gate (hetero dielectric). Using SiGe at the source gives extra degree of freedom to modulate the band gap at source. The limitation of the DMG technique is that as the device dimension scales down, fabrication complexity increases particularly for the hetero dielectric devices. Next we have worked on strain engineering for SiGe source TFET. We have carried out re117 118 search on different device structures and observed that the compressive strain in SiGe, improves the subthreshold swing and ON current as compared to relaxed-SiGe and Si TFETs. As a result of the compressive strain, valence and conduction band split and the effective band gap reduces. Additionally, from pocket implant study it can be concluded that it’s not an effective method to increase the ON current, due to very large increase in OFF current because of electric field enhanced SRH recombination. The source extended structures are also scrutinized to improve the TFET performance. It is observed that 2 nm source extension inside the gate results in the best ON current and subthreshold slope. Length extension inside the gate degrade the sub-threshold slope, but increase the ON current significantly. Additionally, we have used the density functional theory to study thickness dependent study for extended Ge source lateral TFET. We have extended source by 2 nm inside the gate in order to confine the tunneling in Ge. We have observed that the band alignment at Si/Ge junction changes significantly with thickness scaling. It is also found that at reduced thickness, 2 nm overlap length is insufficient to confine the tunneling in Ge. In ultra thin devices, the electrostatic effects also play an important role in BTBT due to increased gate coupling, and this enhances the subsurface BTBT current significantly. The increased contribution of the subsurface BTBT current, to the total drain current in ultra-thin Ge TFET, results in significantly improved subthreshold region characteristics. Following it, we have presented a comparative study of vertical and lateral tunneling in Ge source TFETs, as the source is scaled down from bulk to ultra-thin slab of 4 nm thickness. We have used an overlapped gate over Ge source to include vertical line tunneling along with lateral tunneling in the device. We have observed that as the Ge source thickness is scaled down to ultra-thin films (TGe ≤ 8 nm), line tunneling disappears in the gate-source overlapped TFETs and the net current is degraded. Thus, we find that gate overlapped structure with ultra-thin Ge source is not suitable for future low power applications. We have shown that TFETs with strong lateral P-P tunneling result in better performance at reduced thickness, as compared to the gate-source overlap TFETs, which offer significant line tunneling only in the bulk case. We have also developed a surface potential based analytical model of the drain current, terminal charges, and capacitances for a DG-TFET. The surface potential model includes the 7.1 Scope for Future Work 119 effect of mobile charges and is valid over a wide range of applied bias conditions including both the depletion and the saturation regimes. This model is further used to model the ambipolar current in the device. Additionally, we calculate terminal charge and terminal capacitance. Our model is in good agreement with the TCAD based simulation data for the electric field, drain current, terminal charges, and terminal capacitances in a DG-TFET over a wide operating range. This highlights the fact that the proposed model captures all the relevant aspects of the device physics reasonably well, and it can be very useful as a compact model for computationally efficient simulations for of DG-TFET circuits. 7.1 Scope for Future Work Further research work in the areas, covered in this thesis can be carried out as follows: The effect of strain is incorporated on the bulk properties of SiGe. It would be interesting to include the effect of confinement along with the strain in SiGe and Si as the thickness is scaled. The direct valley of Ge can be engineered by using strain, or by incorporating larger atom like Sn in the lattice structure. Effect of thickness dependent quantum confinement on the valence band of Ge is another area that can be explored. Incorporating the effect of strain on the confined Si/Ge interface is another challenging area for research. In our work, we have solved the 2-D Poisson equation without the mobile charge carrier and have considered the mobile charge effect by empirical equation in the center potential and inverse decay length. It would be interesting to solve the Poisson equation with mobile charge to make the model simpler. Modeling the hetero junction TFET’s with III-V alloys is another area to be explored. The presented model methodology can also be adopted to model the III-V hetero junction TFETs by incorporating the effect of hetero junction. Appendix A List of Publications A.1 Journal Papers • P. Jain, V. Prabhat, and B. Ghosh, “Dual metal-double gate tunnel field effect transistor with mono/hetero dielectric gate material”, accepted in Journal of Computational Electronics, vol. 14, no. 2, pp. 537-542, Jun 2015. • P. Jain, P. Rastogi, C. Yadav, A. Agarwal and Y. S. Chauhan, “Band-to-band tunneling in valley for Ge source lateral tunnel field effect transistor: Thickness scaling ”, Journal of Applied Physics, vol. 122, no. 1, pp.014502, 2017. • P. Jain, C. Yadav, A. Agarwal and Y. S. Chauhan, “Surface potential based modeling of charge, current, and capacitances in DGTFET including mobile channel charge and ambipolar behaviour”, Solid-State Electronics, Vol. 134, no. Supplement C, pp. 74-81, 2017. • B. Tripathi and P. Jain, “SiGe Source Dual Metal Double Gate Tunnel Field Effect Transistor”, Journal of Low Power Electronics, Vol. 13, no. 1, pp. 76-82, 2017. A.2 Conference Papers • P. Jain, P. Rastogi, T. Dutta, A. Agarwal and Y. S. Chauhan, “Impact of Thickness Scaling on Vertical and Lateral Tunneling in Ge Source Tunnel FETs”, IEEE International 120 A.2 Conference Papers 121 Conference on Emerging Electronics (ICEE), Mumbai, India, Dec. 2016. • P. Jain and D. Kumar, “Drive Current Boosting Using Pocket Implant Near to the Strained SiGe/Si Source with Single-Metal/Dual-Metal Double Gate Tunnel Field Effect Transistor”, in Proceedings of International Conference on Intelligent Communication, Control and Devices:ICICCD 2016, Springer Singapore, Oct. 2016. View publication stats