Unit-1. MOS Transistor Theory

Similar documents
Notes. (Subject Code: 7EC5)

Transistor was first invented by William.B.Shockley, Walter Brattain and John Bardeen of Bell Labratories. In 1961, first IC was introduced.

INTRODUCTION TO MOS TECHNOLOGY

420 Intro to VLSI Design

Semiconductor Physics and Devices

CMOS Digital Integrated Circuits Lec 2 Fabrication of MOSFETs

ECE 5745 Complex Digital ASIC Design Topic 2: CMOS Devices

EE4800 CMOS Digital IC Design & Analysis. Lecture 1 Introduction Zhuo Feng

Basic Fabrication Steps

VLSI Design. Introduction

Lecture 0: Introduction

EE301 Electronics I , Fall

FABRICATION OF CMOS INTEGRATED CIRCUITS. Dr. Mohammed M. Farag

VLSI Design. Introduction

EC0306 INTRODUCTION TO VLSI DESIGN

+1 (479)

MOSFET & IC Basics - GATE Problems (Part - I)

EE 5611 Introduction to Microelectronic Technologies Fall Thursday, September 04, 2014 Lecture 02

Solid State Devices- Part- II. Module- IV

Chapter 3: Basics Semiconductor Devices and Processing 2006/9/27 1. Topics

UNIT-VI FIELD EFFECT TRANSISTOR. 1. Explain about the Field Effect Transistor and also mention types of FET s.

Integrated diodes. The forward voltage drop only slightly depends on the forward current. ELEKTRONIKOS ĮTAISAI

In this lecture we will begin a new topic namely the Metal-Oxide-Semiconductor Field Effect Transistor.

Semiconductor Devices

ECE 340 Lecture 37 : Metal- Insulator-Semiconductor FET Class Outline:

Chapter 3 Basics Semiconductor Devices and Processing

UNIT-II LOW POWER VLSI DESIGN APPROACHES

Jack Keil Wolf Lecture. ESE 570: Digital Integrated Circuits and VLSI Fundamentals. Lecture Outline. MOSFET N-Type, P-Type.

Topic 3. CMOS Fabrication Process

INTRODUCTION: Basic operating principle of a MOSFET:

TECHNO INDIA BATANAGAR (DEPARTMENT OF ELECTRONICS & COMMUNICATION ENGINEERING) QUESTION BANK- 2018

ECE520 VLSI Design. Lecture 2: Basic MOS Physics. Payman Zarkesh-Ha

(Refer Slide Time: 02:05)

PAPER SOLUTION_DECEMBER_2014_VLSI_DESIGN_ETRX_SEM_VII Prepared by Girish Gidaye

Power MOSFET Zheng Yang (ERF 3017,

Design Simulation and Analysis of NMOS Characteristics for Varying Oxide Thickness

ESE 570: Digital Integrated Circuits and VLSI Fundamentals

AE53/AC53/AT53/AE103 ELECT. DEVICES & CIRCUITS DEC 2015

4.1 Device Structure and Physical Operation

ECE4902 B2015 HW Set 1

UNIT 3: FIELD EFFECT TRANSISTORS

UNIT 3 Transistors JFET

Layout of a Inverter. Topic 3. CMOS Fabrication Process. The CMOS Process - photolithography (2) The CMOS Process - photolithography (1) v o.

Three Terminal Devices

EE 330 Lecture 7. Design Rules. IC Fabrication Technology Part 1

EMT 251 Introduction to IC Design

UNIT-1 Fundamentals of Low Power VLSI Design

VLSI Technology Dr. Nandita Dasgupta Department of Electrical Engineering Indian Institute of Technology, Madras

CMOS Technology. 1. Why CMOS 2. Qualitative MOSFET model 3. Building a MOSFET 4. CMOS logic gates. Handouts: Lecture Slides. metal ndiff.

3084 IEEE TRANSACTIONS ON NUCLEAR SCIENCE, VOL. 60, NO. 4, AUGUST 2013

PHYSICAL STRUCTURE OF CMOS INTEGRATED CIRCUITS. Dr. Mohammed M. Farag

2.8 - CMOS TECHNOLOGY

Chapter 3 CMOS processing technology (II)

Lecture 020 ECE4430 Review II (1/5/04) Page 020-1

EE 42/100 Lecture 23: CMOS Transistors and Logic Gates. Rev A 4/15/2012 (10:39 AM) Prof. Ali M. Niknejad

Chapter 2 : Semiconductor Materials & Devices (II) Feb

Lecture 020 ECE4430 Review II (1/5/04) Page 020-1

Conduction Characteristics of MOS Transistors (for fixed Vds)! Topic 2. Basic MOS theory & SPICE simulation. MOS Transistor

Topic 2. Basic MOS theory & SPICE simulation

Conduction Characteristics of MOS Transistors (for fixed Vds) Topic 2. Basic MOS theory & SPICE simulation. MOS Transistor

Digital Integrated Circuits A Design Perspective. The Devices. Digital Integrated Circuits 2nd Devices

FIELD EFFECT TRANSISTOR (FET) 1. JUNCTION FIELD EFFECT TRANSISTOR (JFET)

Integrated Circuits: FABRICATION & CHARACTERISTICS - 4. Riju C Issac

Module-3: Metal Oxide Semiconductor (MOS) & Emitter coupled logic (ECL) families

MEASUREMENT AND INSTRUMENTATION STUDY NOTES UNIT-I

Field Effect Transistors (npn)

PROCESS-VOLTAGE-TEMPERATURE (PVT) VARIATIONS AND STATIC TIMING ANALYSIS

Lecture Integrated circuits era

Student Lecture by: Giangiacomo Groppi Joel Cassell Pierre Berthelot September 28 th 2004

VLSI DESIGN AUTOMATION COURSE NOTES THE PRINCIPLES OF VLSI DESIGN

Field-Effect Transistor (FET) is one of the two major transistors; FET derives its name from its working mechanism;

! Review: MOS IV Curves and Switch Model. ! MOS Device Layout. ! Inverter Layout. ! Gate Layout and Stick Diagrams. ! Design Rules. !

MICROPROCESSOR TECHNOLOGY

Organic Electronics. Information: Information: 0331a/ 0442/

UNIT III VLSI CIRCUIT DESIGN PROCESSES. In this chapter we will be studying how to get the schematic into stick diagrams or layouts.

Introduction to VLSI ASIC Design and Technology

FET. Field Effect Transistors ELEKTRONIKA KONTROL. Eka Maulana, ST, MT, M.Eng. Universitas Brawijaya. p + S n n-channel. Gate. Basic structure.

UNIT-III POWER ESTIMATION AND ANALYSIS

! Review: MOS IV Curves and Switch Model. ! MOS Device Layout. ! Inverter Layout. ! Gate Layout and Stick Diagrams. ! Design Rules. !

ESE 570: Digital Integrated Circuits and VLSI Fundamentals

Session 10: Solid State Physics MOSFET

CS/ECE 5710/6710. Composite Layout

Microelectronics, BSc course

Spiral 1 / Unit 8. Transistor Implementations CMOS Logic Gates

NAME: Last First Signature

VLSI Technology Dr. Nandita Dasgupta Department of Electrical Engineering Indian Institute of Technology, Madras

Photolithography I ( Part 1 )

Semiconductor Memory: DRAM and SRAM. Department of Electrical and Computer Engineering, National University of Singapore

A Survey of the Low Power Design Techniques at the Circuit Level

SRM INSTITUTE OF SCIENCE AND TECHNOLOGY (DEEMED UNIVERSITY)

INTERNATIONAL JOURNAL OF APPLIED ENGINEERING RESEARCH, DINDIGUL Volume 1, No 3, 2010

Lecture 2 p-n junction Diode characteristics. By Asst. Prof Dr. Jassim K. Hmood

Laboratory #5 BJT Basics and MOSFET Basics

EC 1354-Principles of VLSI Design

EECS 151/251A Spring 2019 Digital Design and Integrated Circuits. Instructors: Wawrzynek. Lecture 8 EE141

Lecture-45. MOS Field-Effect-Transistors Threshold voltage

An introduction to Depletion-mode MOSFETs By Linden Harrison

Department of Electrical Engineering IIT Madras

Pramoda N V Department of Electronics and Communication Engineering, MCE Hassan Karnataka India

Advantage of Having Large Numbers of Function on a Single Chip. Less Area occupied Less power Consumption Higher Speed Higher Reliability Economical

Transcription:

VLSI DESIGN -EEE

Unit-1 MOS Transistor Theory

VLSI DESIGN UNIT I Contents: 1.1 Historical Perspective 1.2 What is VLSI? - Introduction 1.3 VLSI Design Flow 1.4 Design Hierarchy 1.5 Basic MOS Transistor 1.6 CMOS Chip Fabrication 1.7 Layout Design Rules 1.8 Lambda Based Rules 1.9 Design Rules - MOSIS Scalable CMOS (SCMOS)

Objective: * To show the evolution of logic complexity in integrated circuits. * To understand what is VLSI? * To illustrate a design flow for logic chips using Y-chart. * To understand the divide and conquer technique of dividing a module into submodules for the simplicity of design. * To know the structure, symbol and operation of basic MOS Transistor. * To know the process flow in chip fabrication and the interaction of various processing steps. * To have an overview about Advanced CMOS fabrication technologies. * To specify the layout design rules in two ways i) Lambda rules ii)micron rules.

1.1 Historical Perspective The electronics industry has achieved a phenomenal growth over the last two decades, mainly due to the rapid advances in integration technologies, large-scale systems design - in short, due to the advent of VLSI. The number of applications of integrated circuits in high-performance computing, telecommunications, and consumer electronics has been rising steadily, and at a very fast pace. Typically, the required computational power (or, in other words, the intelligence) of these applications is the driving force for the fast development of this field. Figure 1.1 gives an overview of the prominent trends in information technologies over the next few decades. The current leading-edge technologies (such as low bit-rate video and cellular communications) already provide the end-users a certain amount of processing power and portability. This trend is expected to continue, with very important implications on VLSI and systems design. One of the most important characteristics of information services is their increasing need for very high processing power and bandwidth (in order to handle real-time video, for example). The other important characteristic is that the information services tend to become more and more personalized (as opposed to collective services such as broadcasting), which means that the devices must be more intelligent to answer individual demands, and at the same time they must be portable to allow more flexibility/mobility. Figure-1.1: Prominent trends in information service technologies. As more and more complex functions are required in various data processing and telecommunications devices, the need to integrate these functions in a small system/package is also increasing. The level of integration as measured by the number of logic gates in a monolithic chip has been steadily rising for almost three decades, mainly

due to the rapid progress in processing technology and interconnect technology. Table 1.1 shows the evolution of logic complexity in integrated circuits over the last three decades, and marks the milestones of each era. Here, the numbers for circuit complexity should be interpreted only as representative examples to show the order-of-magnitude. A logic block can contain anywhere from 10 to 100 transistors, depending on the function. Stateof-the-art examples of ULSI chips, such as the DEC Alpha or the INTEL Pentium contain 3 to 6 million transistors. ERA (number of logic blocks per chip) Single transistor Unit logic (one gate) Multi-function Complex function Medium Scale Integration Large Scale Integration (LSI) Very Large Scale Integration (VLSI) Ultra Large Scale Integration DATE COMPLEXITY 1959 1960 1962 1964 1967 1972 less than 1 1 2-4 5-20 20-200 (MSI) 200-2000 1978 2000-20000 1989 20000 -? (ULSI) Table-1.1: Evolution of logic complexity in integrated circuits. The most important message here is that the logic complexity per chip has been (and still is) increasing exponentially. The monolithic integration of a large number of functions on a single chip usually provides: Less area/volume and therefore, compactness Less power consumption Less testing requirements at system level Higher reliability, mainly due to improved on-chip interconnects Higher speed, due to significantly reduced interconnection length Significant cost savings

Figure-1.2: Evolution of integration density and minimum feature size, as seen in the early 1980s. Therefore, the current trend of integration will also continue in the foreseeable future. Advances in device manufacturing technology, and especially the steady reduction of minimum feature size (minimum length of a transistor or an interconnect realizable on chip) support this trend. Figure 1.2 shows the history and forecast of chip complexity - and minimum feature size - over time, as seen in the early 1980s. At that time, a minimum feature size of 0.3 microns was expected around the year 2000. The actual development of the technology, however, has far exceeded these expectations. A minimum size of 0.25 microns was readily achievable by the year 1995. As a direct result of this, the integration density has also exceeded previous expectations - the first 64 Mbit DRAM, and the INTEL Pentium microprocessor chip containing more than 3 million transistors were already available by 1994, pushing the envelope of integration density. When comparing the integration density of integrated circuits, a clear distinction must be made between the memory chips and logic chips. Figure 1.3 shows the level of integration over time for memory and logic chips, starting in 1970. It can be observed that in terms of transistor count, logic chips contain significantly fewer transistors in any given year mainly due to large consumption of chip area for complex interconnects. Memory circuits are highly regular and thus more cells can be integrated with much less area for interconnects.

Figure-1.3: Level of integration over time, for memory chips and logic chips. Generally speaking, logic chips such as microprocessor chips and digital signal processing (DSP) chips contain not only large arrays of memory (SRAM) cells, but also many different functional units. As a result, their design complexity is considered much higher than that of memory chips, although advanced memory chips contain some sophisticated logic functions. The design complexity of logic chips increases almost exponentially with the number of transistors to be integrated. This is translated into the increase in the design cycle time, which is the time period from the start of the chip development until the masktape delivery time. However, in order to make the best use of the current technology, the chip development time has to be short enough to allow the maturing of chip manufacturing and timely delivery to customers. As a result, the level of actual logic integration tends to fall short of the integration level achievable with the current processing technology. Sophisticated computer-aided design (CAD) tools and methodologies are developed and applied in order to manage the rapidly increasing design complexity.

1.2 INTRODUCTION What is VLSI? VLSI stands for "Very Large Scale Integration". This is the field which involves packing more and more logic devices into smaller and smaller areas. Thanks to VLSI, circuits that would have taken boardfuls of space can now be put into a small space few millimeters across! This has opened up a big opportunity to do things that were not possible before. VLSI circuits are everywhere... your computer, your car, your brand new state-of-the-art digital camera, the cell-phones, and what have you. All this involves a lot of expertise on many fronts within the same field. VLSI has been around for a long time, there is nothing new about it... but as a side effect of advances in the world of computers, there has been a dramatic proliferation of tools that can be used to design VLSI circuits. Alongside, obeying Moore's law, the capability of an IC has increased exponentially over the years, in terms of computation power, utilization of available area, yield. The combined effect of these two advances is that people can now put diverse functionality into the IC's, opening up new frontiers. Examples are embedded systems, where intelligent devices are put inside everyday objects, and ubiquitous computing where small computing devices proliferate to such an extent that even the shoes you wear may actually do something useful like monitoring your heartbeats! DEALING WITH VLSI CIRCUITS Digital VLSI circuits are predominantly CMOS based. The way normal blocks like latches and gates are implemented is different from what students have seen so far, but the behavior remains the same. All the miniaturization involves new things to consider. A lot of thought has to go into actual implementations as well as design. Let us look at some of the factors involved... 1. Circuit Delays Large complicated circuits running at very high frequencies have one big problem to tackle - the problem of delays in propagation of signals through gates and wires... even for areas a few micrometers across! The operation speed is so large that as the delays add up, they can actually become comparable to the clock speeds. 2. Power. Another effect of high operation frequencies is increased consumption of power. This has two-fold effect - devices consume batteries faster, and heat dissipation increases. Coupled with the fact that surface areas have decreased, heat poses a major threat to the stability of the circuit itself.

3. Layout. Laying out the circuit components is task common to all branches of electronics. Whats so special in our case is that there are many possible ways to do this; there can be multiple layers of different materials on the same silicon, there can be different arrangements of the smaller parts for the same component and so on. The power dissipation and speed in a circuit present a trade-off; if we try to optimise on one, the other is affected. The choice between the two is determined by the way we chose the layout the circuit components. Layout can also affect the fabrication of VLSI chips, making it either easy or difficult to implement the components on the silicon. 1.3 VLSI Design Flow The design process, at various levels, is usually evolutionary in nature. It starts with a given set of requirements. Initial design is developed and tested against the requirements. When requirements are not met, the design has to be improved. If such improvement is either not possible or too costly, then the revision of requirements and its impact analysis must be considered. The Y-chart (first introduced by D. Gajski) shown in Fig. 1.4 illustrates a design flow for most logic chips, using design activities on three different axes (domains) which resemble the letter Y.

Figure-1.4: Typical VLSI design flow in three domains (Y-chart representation). The Y-chart consists of three major domains, namely: behavioral domain, structural domain, geometrical layout domain. The design flow starts from the algorithm that describes the behavior of the target chip. The corresponding architecture of the processor is first defined. It is mapped onto the chip surface by floorplanning. The next design evolution in the behavioral domain defines finite state machines (FSMs) which are structurally implemented with functional modules such as registers and arithmetic logic units (ALUs). These modules are then geometrically placed onto the chip surface using CAD tools for automatic module placement followed by routing, with a goal of minimizing the interconnects area and signal delays. The third evolution starts with a behavioral module description. Individual modules are then implemented with leaf cells. At this stage the chip is described in terms of logic gates (leaf cells), which can be placed and interconnected by using a cell placement & routing program. The last evolution involves a detailed Boolean description of leaf cells followed by a transistor level implementation of leaf cells and mask generation. In standard-cell based design, leaf cells are already pre-designed and stored in a library for logic design use.

Figure-1.5: A more simplified view of VLSI design flow. Figure 1.5 provides a more simplified view of the VLSI design flow, taking into account the various representations, or abstractions of design - behavioral, logic, circuit and mask layout. Note that the verification of design plays a very important role in every step during this process. The failure to properly verify a design in its early phases typically causes significant and expensive re-design at a later stage, which ultimately increases the timeto-market. Although the design process has been described in linear fashion for simplicity, in reality there are many iterations back and forth, especially between any two neighboring steps, and occasionally even remotely separated pairs. Although top-down design flow provides an excellent design process control, in reality, there is no truly unidirectional top-down design flow. Both top-down and bottom-up approaches have to be combined. For instance, if a chip designer defined an architecture without close estimation of the

corresponding chip area, then it is very likely that the resulting chip layout exceeds the area limit of the available technology. In such a case, in order to fit the architecture into the allowable chip area, some functions may have to be removed and the design process must be repeated. Such changes may require significant modification of the original requirements. Thus, it is very important to feed forward low-level information to higher levels (bottom up) as early as possible. In the following, we will examine design methodologies and structured approaches which have been developed over the years to deal with both complex hardware and software projects. Regardless of the actual size of the project, the basic principles of structured design will improve the prospects of success. Some of the classical techniques for reducing the complexity of IC design are: Hierarchy, regularity, modularity and locality. 1.4 Design Hierarchy The use of hierarchy, or divide and conquer technique involves dividing a module into sub- modules and then repeating this operation on the sub-modules until the complexity of the smaller parts becomes manageable. This approach is very similar to the software case where large programs are split into smaller and smaller sections until simple subroutines, with well-defined functions and interfaces, can be written. In Section 1.2, we have seen that the design of a VLSI chip can be represented in three domains. Correspondingly, a hierarchy structure can be described in each domain separately. However, it is important for the simplicity of design that the hierarchies in different domains can be mapped into each other easily. As an example of structural hierarchy, Fig. 1.6 shows the structural decomposition of a CMOS four-bit adder into its components. The adder can be decomposed progressively into one- bit adders, separate carry and sum circuits, and finally, into individual logic gates. At this lower level of the hierarchy, the design of a simple circuit realizing a welldefined Boolean function is much more easier to handle than at the higher levels of the hierarchy. In the physical domain, partitioning a complex system into its various functional blocks will provide a valuable guidance for the actual realization of these blocks on chip. Obviously, the approximate shape and size (area) of each sub-module should be estimated in order to provide a useful floor plan. Figure 1.7 shows the hierarchical decomposition of a four-bit adder in physical description (geometrical layout) domain, resulting in a simple floor plan. This physical view describes the external geometry of the adder, the locations of input and output pins, and how pin locations allow some signals (in this case the carry signals) to be transferred from one sub-block to the other without external routing. At lower levels of the physical hierarchy, the internal mask

Figure-1.6: Structural decomposition of a four-bit adder circuit, showing the hierarchy down to gate level. 1.5 Basic MOS Transistor The most basic element in the design of a large scale integrated circuit is the transistor. Metal-Oxide-Semiconductor Field Effect Transistors (MOSFET) are formed as a sandwich consisting of a semiconductor layer, usually a slice, or wafer, from a single crystal of silicon; a layer of silicon dioxide (the oxide) and a layer of metal. These layers are patterned in a manner which permits transistors to be formed in the semiconductor material (the substrate ); a diagram showing a typical (idealized) MOSFET is shown in Figure Silicon dioxide is a very good insulator, so a very thin layer, typically only a few hundred molecules thick, is required. Actually, the transistors which we will use do not use metal for their gate regions, but instead use polycrystalline silicon (poly). Polysilicon gate FET's have replaced virtually all of the older devices using metal gates in large scale integrated circuits. (Both metal and polysilicon FET's are sometimes referred to as IGFET's --- insulated gate field effect transistors, since the silicon dioxide under the gate is an insulator. Figure: MOS transistor

The transistor consists of three regions, labeled the source', the gate and the ``drain''. The area labeled as the gate region is actually a ``sandwich'' consisting of the underlying substrate material, which is a single crystal of semiconductor material (usually silicon); a thin insulating layer (usually silicon dioxide); and an upper metal layer. Electrical charge, or current, can flow from the source to the drain depending on the charge applied to the gate region. The semiconductor material in the source and drain region are ``doped'' with a different type of material than in the region under the gate, so an NPN or PNP type structure exists between the source and drain region of a MOSFET. An MOS transistor is a majority-carrier device, in which the current in a conducting channel between the source and the drain is modulated by a voltage applied to the gate. Symbols NMOS (n-type MOS transistor) (1) Majority carrier = electrons (2) A positive voltage applied on the gate with respect to the substrate enhances the number of electrons in the channel and hence increases the conductivity of the channel. (3) If gate voltage is less than a threshold voltage Vt, the channel is cut-off (very low current between source & drain). PMOS (p-type MOS transistor) (1) Majority carrier = holes (2) Applied voltage is negative with respect to substrate. Threshold voltage (Vt): The voltage at which an MOS device begins to conduct ("turn on") Relationship between Vgs (gate-to-source voltage) and the source-to-drain current (Ids), given a fixed drain-to-source voltage (Vds).

MOS Transistor types based on operation: (1) Devices that are normally cut-off with zero gate bias are classified as "enhancementmode "devices. (2) Devices that conduct with zero gate bias are called "depletion-mode "devices. (3) Enhancement-mode devices are more popular in practical use. 1.4.1 NMOS Enhancement Transistor Consist of (1) Moderately doped p-type silicon substrate (2) Two heavily doped n + regions, the source and drain, are diffused. (3) Channel is covered by a thin insulating layer of silicon dioxide (SiO2) called " Gate Oxide " (4) Over the oxide is a polycrystalline silicon (polysilicon) electrode, referred to as the "Gate". Features (1) Since the oxide layer is an insulator, the DC current from the gate to channel is essentially zero. (2) No physical distinction between the drain and source regions. (3) Since SiO2 has low loss and high dielectric strength, the application of high gate fields is feasible.

In operation (1) Set Vds > 0 in operation (2) Vgs =0 no current flow between source and drain. They are insulated by two reversed-biased PN junctions (3) When Vg > 0, the produced E field attracts electrons toward the gate and repels holes. (4) If Vg is sufficiently large, the region under the gate changes from p-type to ntype(due to accumulation of attracted elections) and provides a conducting path between source and drain. >The thin layer of p-type silicon is said to be "inverted". (5) Three modes a. Accumulation mode (Vgs << Vt) b. Depletion mode (Vgs =Vt) c. Inversion mode (Vgs > Vt)

Electrically (1) An MOS device can be considered as a voltage-controlled switch that conducts when Vgs >Vt (given Vds>0) (2) An MOS device can be considered as a voltage-controlled resistor Effective gate voltage (Vgs-Vt) At the source end, the full gate voltage is effective in inverting the channel. At the drain end, only the difference between the gate and drain voltage is effective.

1.4.2 PMOS Enhancement Transistor (1) Vg < 0 (2) Holes are major carrier (3) Vd < 0, which sweeps holes from the source through the channel to the drain. Current Voltage curves:

1.4.3 Threshold voltage The threshold voltage of a MOSFET is usually defined as the gate voltage where an inversion layer forms at the interface between the insulating layer (oxide) and the substrate (body) of the transistor. The creation of this layer is described next. In an n-mosfet the substrate of the transistor is composed of p-type silicon (see doping (semiconductor)), which has positively charged mobile holes as carriers. When a positive voltage is applied on the gate, an electric field causes the holes to be repelled from the interface, creating a depletion region containing immobile negatively charged acceptor ions. A further increase in the gate voltage eventually causes electrons to appear at the interface, in what is called an inversion layer, or channel. Historically the gate voltage at which the electron density at the interface is the same as the hole density in the neutral bulk material is called the threshold voltage. Practically speaking the threshold voltage is the voltage at which there are sufficient electrons in the inversion layer to make a low resistance conducting path between the MOSFET source and drain. It is a function of (1) Gate conductor material (2) Gate insulator material (3) Gate insulator thickness (4) Impurity at the silicon-insulator interface (5) Voltage between the source and the substrate Vsb (6) Temperature a. -4 mv/ C high substrate doping b. -2 mv/ C low substrate doping MOS Transistor structure The basic concept of a MOS Transistor is simple and best understood by looking at its structure:

It is always an integrated structure, there are practically no single individual MOS transistors. A MOS transistor is primarily a switch for digital devices. Ideally, it works as follows: If the voltage at the gate electrode is "on", the transistor is "on", too, and current flow between the source and drain electrodes is possible (almost) without losses. If the voltage at the gate electrode is "off", the transistor is "off", too, and no current flows between the source and drain electrode. In reality, this only works for a given polarity of the gate voltage. Moreover, a MOS transistor needs very thin gate dielectrics (around, or better below 10 nm), and extreme control of materials and technologies if real MOS transistors are to behave as they are expected to in "ideal" theory. Understanding MOS transistor qualitatively is easy. We look at the example from above and apply some source-drain voltage VSD in either polarity, but no gate voltage yet. What we have under these conditions is: A n-type Si substrate with a certain equilibrium density of electrons ne(ug = 0), or ne(0) for short. Its value is entirely determined by doping (and the temperature, which we will neglect at the present, however) and is the same everywhere. We also have a much smaller concentration nh(0) of holes. Two pn-junctions, one of which is polarized in forward direction (the one with the positive voltage pole), and the other one in reverse. This is true for any polarity; in particular one junction will always be biased in reverse. Therefore no source-drain current ISD will flow (or only some small reverse current which we will neglect at present). There will also be no current in the forwardly biased diode, because the n-si of the substrate in the figure is not electrically connected to anything (in reality, we might simply ground the positive USD pole and the substrate). For a gate voltage UG = 0 V, there are no currents and everything is in equilibrium. But now apply a negative voltage at the gate. The electrons in the substrate below the gate will be electrostatically repelled and driven into the substrate. Their concentration directly below the gate will go down, ne (U) will be a function of the depth coordinate z. Since we still have equilibrium, the mass action law for carriers holds anywhere in the Si, i.e.. ne (z) nh (z) = ni2 With ni = intrinsic carrier density in Si This gives us nh (z) = ni2 ne (z) In other words: If the electron concentration below the gate goes down, the hole concentration goes up. If we sufficiently decrease the electron concentration under the gate by cranking up the gate voltage, we will eventually achieve the condition nh (z = 0) = ne (z = 0) right under the gate, i.e. at z = 0. If we increase the gate voltage even more, we will encounter the

condition nh (z) > ne (z) for small values of z, ie. for zc > z > 0. In other words: Right under the gate we now have more holes than electrons; this is called a state of inversion for obvious reasons. Si having more holes than electrons is also called p-type Si. What we have now is a p-conducting channel (with width zc) connecting the p-conducting source and drain. There are no more pn-junctions preventing current flow under the gate current can flow freely; only limited by the ohmic resistance of contacts, source/drain and channel. The resistivity of this channel will be determined by the amount of Si we have inverted; it will rapidly come down with the voltage as soon as the threshold voltage necessary for inversion is reached. If we reverse the voltage at the gate, we attract electrons and their concentration under the gate increases. This is called a state of accumulation. The pn junctions at source and drain stay intact, and no source - drain current will flow. Obviously, if we want to switch a MOS transistor "on" with a positive gate voltage, we must now reverse the doping and use a p-doped substrates and n-doped source/drain regions. The two basic types we call "n-channel MOS" and "p-channel MOS" according to the kind of doping in the channel upon inversion (or the source/drain contacts).

Voltage at the gate Conditions in the Si Voltage drop Charge distribution Nothing happens. The band in the substrate is perfectly flat (and so is the band in the contact electrode, but that is of no interest). We only There are no would have a net charges voltage (or better potential) drop, if the Fermi energies of substrate and gate electrode were different Zero gate voltage. "Flat band" condition Positive gate voltage. Accumulation With a positive The voltage voltage at the drops mostly gate we attract in the oxide the electrons in the substrate. The bands must bend down somewhat, and we increase the number of electrons in the conduction band accordingly. (There is a bit of a space charge region (SCR) in the contact, but There is some positive charge at the gate electrode interface (with our Si electrode from the SCR), and negative charge from the many electrons in the (thin) accumulation layer on the other side of the gate dielectric.

that is of no interest). Small negative gate voltage. Depletion With a (small) negative voltage at the gate, we repel the electrons in the substrate. Their concentration decreases, the hole concentration is still low - we have a layer depleted of mobile carriers and therefore a SCR. The voltage drops mostly in the oxide, but also to some extent in the SCR. There is some negative charge at the gate electrode interface (accumulated electrons with our Si electrode), and positive charge smeared out in the the (extended) SCR layer on the other side of the gate dielectric. With a (large) negative voltage at the gate, we repel the electrons in the substrate very much. The bands bend so much, that the Fermi energy (red line) is in the lower half of the band close to the interface. In this region holes are the The voltage drops mostly in the oxide, but also to some extent in the SCR and the inversion layer. There is more negative charge at the gate electrode interface (accumulated electrons with our Si electrode), some positive charge smeared out in the the (extended) SCR layer on the other side of the gate dielectric, and Large negagive gate voltage. Inversion

majority carriers, we gave inversion. We still have a SCR, too. a lot of positive charge from the holes in thin inversion layer. 1.6 CMOS CHIP FABRICATION Contents : 1.5.1 Introduction 1.5.2 Fabrication Process Flow - Basic Steps 1.5.3 The CMOS n-well Process 1.5.4 Advanced CMOS Fabrication Technologies 1.5.1 Introduction In this topic, the emphasis will be on the general outline of the process flow and on the interaction of various processing steps, which ultimately determine the device and the circuit performance characteristics. The following chapters show that there are very strong links between the fabrication process, the circuit design process and the performance of the resulting chip. Hence, circuit designers must have a working knowledge of chip fabrication to create effective designs and in order to optimize the circuits with respect to various manufacturing parameters. Also, the circuit designer must have a clear understanding of the roles of various masks used in the fabrication process, and how the masks are used to define various features of the devices on-chip.

The following discussion will concentrate on the well-established CMOS fabrication technology, which requires that both n-channel (nmos) and p-channel (pmos) transistors be built on the same chip substrate. To accommodate both nmos and pmos devices, special regions must be created in which the semiconductor type is opposite to the substrate type. These regions are called wells or tubs. A p-well is created in an n-type substrate or, alternatively, an n- well is created in a p-type substrate. In the simple n-well CMOS fabrication technology presented, the nmos transistor is created in the p-type substrate, and the pmos transistor is created in the n-well, which is built-in into the ptype substrate. In the twin-tub CMOS technology, additional tubs of the same type as the substrate can also be created for device optimization. The simplified process sequence for the fabrication of CMOS integrated circuits on a ptype silicon substrate is shown in Fig. 2.1. The process starts with the creation of the nwell regions for pmos transistors, by impurity implantation into the substrate. Then, a thick oxide is grown in the regions surrounding the nmos and pmos active regions. The thin gate oxide is subsequently grown on the surface through thermal oxidation. These steps are followed by the creation of n+ and p+ regions (source, drain and channel-stop implants) and by final metallization (creation of metal interconnects). Figure-2.1: Simplified process sequence for fabrication of the n-well CMOS integrated circuit with a single polysilicon layer, showing only major fabrication steps.

The process flow sequence pictured in Fig. 2.1 may at first seem to be too abstract, since detailed fabrication steps are not shown. To obtain a better understanding of the issues involved in the semiconductor fabrication process, we first have to consider some of the basic steps in more detail. 1.5.2 Fabrication Process Flow - Basic Steps Note that each processing step requires that certain areas are defined on chip by appropriate masks. Consequently, the integrated circuit may be viewed as a set of patterned layers of doped silicon, polysilicon, metal and insulating silicon dioxide. In general, a layer must be patterned before the next layer of material is applied on chip. The process used to transfer a pattern to a layer on the chip is called lithography. Since each layer has its own distinct patterning requirements, the lithographic sequence must be repeated for every layer, using a different mask. To illustrate the fabrication steps involved in patterning silicon dioxide through optical lithography, let us first examine the process flow shown in Fig. 2.2. The sequence starts with the thermal oxidation of the silicon surface, by which an oxide layer of about 1 micrometer thickness, for example, is created on the substrate (Fig. 2.2(b)). The entire oxide surface is then covered with a layer of photoresist, which is essentially a lightsensitive, acid-resistant organic polymer, initially insoluble in the developing solution (Fig. 2.2(c)). If the photoresist material is exposed to ultraviolet (UV) light, the exposed areas become soluble so that the they are no longer resistant to etching solvents. To selectively expose the photoresist, we have to cover some of the areas on the surface with a mask during exposure. Thus, when the structure with the mask on top is exposed to UV light, areas which are covered by the opaque features on the mask are shielded. In the areas where the UV light can pass through, on the other hand, the photoresist is exposed and becomes soluble (Fig. 2.2(d)).

Figure-2.2: Process steps required for patterning of silicon dioxide. The type of photoresist which is initially insoluble and becomes soluble after exposure to UV light is called positive photoresist. The process sequence shown in Fig. 2.2 uses positive photoresist. There is another type of photoresist which is initially soluble and becomes insoluble (hardened) after exposure to UV light, called negative photoresist. If negative photoresist is used in the photolithography process, the areas which are not shielded from the UV light by the opaque mask features become insoluble, whereas the shielded areas can subsequently be etched away by a developing solution. Negative photoresists are more sensitive to light, but their photolithographic resolution is not as high as that of the positive photoresists. Therefore, negative photoresists are used less commonly in the manufacturing of high-density integrated circuits. Following the UV exposure step, the unexposed portions of the photoresist can be removed by a solvent. Now, the silicon dioxide regions which are not covered by hardened photoresist can be etched away either by using a chemical solvent (HF acid) or by using a dry etch (plasma etch) process (Fig. 2.2(e)). Note that at the end of this step, we obtain an oxide window that reaches down to the silicon surface (Fig. 2.2(f)). The remaining photoresist can now be stripped from the silicon dioxide surface by using

another solvent, leaving the patterned silicon dioxide feature on the surface as shown in Fig. 2.2(g). The sequence of process steps illustrated in detail in Fig. 2.2 actually accomplishes a single pattern transfer onto the silicon dioxide surface, as shown in Fig. 2.3. The fabrication of semiconductor devices requires several such pattern transfers to be performed on silicon dioxide, polysilicon, and metal. The basic patterning process used in all fabrication steps, however, is quite similar to the one shown in Fig. 2.2. Also note that for accurate generation of high-density patterns required in sub-micron devices, electron beam (Ebeam) lithography is used instead of optical lithography. In the following, the main processing steps involved in the fabrication of an n-channel MOS transistor on p-type silicon substrate will be examined. Figure-2.3: The result of a single lithographic patterning sequence on silicon dioxide, without showing the intermediate steps. Compare the unpatterned structure (top) and the patterned structure (bottom) with Fig. 2.2(b) and Fig. 2.2(g), respectively. The process starts with the oxidation of the silicon substrate (Fig. 2.4(a)), in which a relatively thick silicon dioxide layer, also called field oxide, is created on the surface (Fig. 2.4(b)). Then, the field oxide is selectively etched to expose the silicon surface on which the MOS transistor will be created (Fig. 2.4(c)). Following this step, the surface is covered with a thin, high-quality oxide layer, which will eventually form the gate oxide of the MOS transistor (Fig. 2.4(d)). On top of the thin oxide, a layer of polysilicon (polycrystalline silicon) is deposited (Fig. 2.4(e)). Polysilicon is used both as gate electrode material for MOS transistors and also as an interconnect medium in silicon integrated circuits. Undoped polysilicon has relatively high resistivity. The resistivity of polysilicon can be reduced, however, by doping it with impurity atoms.

After deposition, the polysilicon layer is patterned and etched to form the interconnects and the MOS transistor gates (Fig. 2.4(f)). The thin gate oxide not covered by polysilicon is also etched away, which exposes the bare silicon surface on which the source and drain junctions are to be formed (Fig. 2.4(g)). The entire silicon surface is then doped with a high concentration of impurities, either through diffusion or ion implantation (in this case with donor atoms to produce n-type doping). Figure 2.4(h) shows that the doping penetrates the exposed areas on the silicon surface, ultimately creating two n-type regions (source and drain junctions) in the p-type substrate. The impurity doping also penetrates the polysilicon on the surface, reducing its resistivity. Note that the polysilicon gate, which is patterned before doping actually defines the precise location of the channel region and, hence, the location of the source and the drain regions. Since this procedure allows very precise positioning of the two regions relative to the gate, it is also called the self-aligned process.

Figure-2.4: Process flow for the fabrication of an n-type MOSFET on p-type silicon.

Once the source and drain regions are completed, the entire surface is again covered with an insulating layer of silicon dioxide (Fig. 2.4(i)). The insulating oxide layer is then patterned in order to provide contact windows for the drain and source junctions (Fig. 2.4(j)). The surface is covered with evaporated aluminum which will form the interconnects (Fig. 2.4(k)). Finally, the metal layer is patterned and etched, completing the interconnection of the MOS transistors on the surface (Fig. 2.4(l)). Usually, a second (and third) layer of metallic interconnect can also be added on top of this structure by creating another insulating oxide layer, cutting contact (via) holes, depositing, and patterning the metal. 1.5.3 The CMOS n-well Process Having examined the basic process steps for pattern transfer through lithography, and having gone through the fabrication procedure of a single n-type MOS transistor, we can now return to the generalized fabrication sequence of n-well CMOS integrated circuits, as shown in Fig. 2.1. In the following figures, some of the important process steps involved in the fabrication of a CMOS inverter will be shown by a top view of the lithographic masks and a cross-sectional view of the relevant areas. The n-well CMOS process starts with a moderately doped (with impurity concentration typically less than 1015 cm-3) p-type silicon substrate. Then, an initial oxide layer is grown on the entire surface. The first lithographic mask defines the n-well region. Donor atoms, usually phosphorus, are implanted through this window in the oxide. Once the n-well is created, the active areas of the nmos and pmos transistors can be defined. Figures 2.5 through 2.10 illustrate the significant milestones that occur during the fabrication process of a CMOS inverter.

Figure-2.5: Following the creation of the n-well region, a thick field oxide is grown in the areas surrounding the transistor active regions, and a thin gate oxide is grown on top of the active regions. The thickness and the quality of the gate oxide are two of the most critical fabrication parameters, since they strongly affect the operational characteristics of the MOS transistor, as well as its long-term reliability.

Figure-2.6: The polysilicon layer is deposited using chemical vapor deposition (CVD) and patterned by dry (plasma) etching. The created polysilicon lines will function as the gate electrodes of the nmos and the pmos transistors and their interconnects. Also, the polysilicon gates act as self-aligned masks for the source and drain implantations that follow this step.

Figure-2.7: Using a set of two masks, the n+ and p+ regions are implanted into the substrate and into the n- well, respectively. Also, the ohmic contacts to the substrate and to the n-well are implanted in this process step.

Figure-2.8: An insulating silicon dioxide layer is deposited over the entire wafer using CVD. Then, the contacts are defined and etched away to expose the silicon or polysilicon contact windows. These contact windows are necessary to complete the circuit interconnections using the metal layer, which is patterned in the next step.

Figure-2.9: Metal (aluminum) is deposited over the entire chip surface using metal evaporation, and the metal lines are patterned through etching. Since the wafer surface is non-planar, the quality and the integrity of the metal lines created in this step are very critical and are ultimately essential for circuit reliability.

Figure-2.10: The composite layout and the resulting cross-sectional view of the chip, showing one nmos and one pmos transistor (built-in n-well), the polysilicon and metal interconnections. The final step is to deposit the passivation layer (for protection) over the chip, except for wire-bonding pad areas. The patterning process by the use of a succession of masks and process steps is conceptually summarized in Fig. 2.11. It is seen that a series of masking steps must be sequentially performed for the desired patterns to be created on the wafer surface. An example of the end result of this sequence is shown as a cross-section on the right.

1.5.4 Advanced CMOS Fabrication Technologies In this section, two examples will be given for advanced CMOS processes which offer additional benefits in terms of device performance and integration density. These processes, namely, the twin-tub CMOS process and the silicon-on-insulator (SOI) process, are becoming especially more popular for sub-micron geometries where device performance and density must be pushed beyond the limits of the conventional n-well CMOS process. Twin-Tub (Twin-Well) CMOS Process This technology provides the basis for separate optimization of the nmos and pmos transistors, thus making it possible for threshold voltage, body effect and the channel transconductance of both types of transistors to be tuned independently. Generally, the starting material is a n+ or p+ substrate, with a lightly doped epitaxial layer on top. This epitaxial layer provides the actual substrate on which the n-well and the p-well are formed. Since two independent doping steps are performed for the creation of the well regions, the dopant concentrations can be carefully optimized to produce the desired device characteristics. In the conventional n-well CMOS process, the doping density of the well region is typically about one order of magnitude higher than the substrate, which, among other effects, results in unbalanced drain parasitics. The twin-tub process (Fig. 2.12) also avoids this problem. Figure-2.12: Cross-section of nmos and pmos transistors in twin-tub CMOS process. Silicon-on-Insulator (SOI) CMOS Process Rather than using silicon as the substrate material, technologists have sought to use an insulating substrate to improve process characteristics such as speed and latch-up susceptibility. The SOI CMOS technology allows the creation of independent, completely isolated nmos and pmos transistors virtually side-by-side on an insulating substrate (for example: sapphire). The main advantages of this technology are the higher integration

density (because of the absence of well regions), complete avoidance of the latch-up problem, and lower parasitic capacitances compared to the conventional n-well or twintub CMOS processes. A cross-section of nmos and pmos devices in created using SOI process is shown in Fig. 2.13. The SOI CMOS process is considerably more costly than the standard n-well CMOS process. Yet the improvements of device performance and the absence of latch-up problems can justify its use, especially for deep-sub-micron devices. Figure-2.13: Cross-section of nmos and pmos transistors in SOI CMOS process. 1.7 Layout Design Rules The physical mask layout of any circuit to be manufactured using a particular process must conform to a set of geometric constraints or rules, which are generally called layout design rules. These rules usually specify the minimum allowable line widths for physical objects on-chip such as metal and polysilicon interconnects or diffusion areas, minimum feature dimensions, and minimum allowable separations between two such features. If a metal line width is made too small, for example, it is possible for the line to break during the fabrication process or afterwards, resulting in an open circuit. If two lines are placed too close to each other in the layout, they may form an unwanted short circuit by merging during or after the fabrication process. The main objective of design rules is to achieve a high overall yield and reliability while using the smallest possible silicon area, for any circuit to be manufactured with a particular process. Note that there is usually a trade-off between higher yield which is obtained through conservative geometries, and better area efficiency, which is obtained through aggressive, high- density placement of various features on the chip. The layout design rules which are specified for a particular fabrication process normally represent a reasonable optimum point in terms of yield and density. It must be emphasized, however, that the design rules do not represent strict boundaries which separate "correct" designs from "incorrect" ones. A layout which violates some of the specified design rules may still result in an operational circuit with reasonable yield, whereas another layout observing all specified design rules may result in a circuit which is not functional and/or has very low yield. To summarize, we can say, in general, that observing the layout design rules significantly increases the probability of fabricating a successful product with high yield.

The design rules are usually described in two ways : Micron rules, in which the layout constraints such as minimum feature sizes and minimum allowable feature separations, are stated in terms of absolute dimensions in micrometers, or, Lambda rules, which specify the layout constraints in terms of a single parameter (?) and, thus, allow linear, proportional scaling of all geometrical constraints. 1.8 Lambda Based Rules Lambda-based layout design rules were originally devised to simplify the industrystandard micron-based design rules and to allow scaling capability for various processes. It must be emphasized, however, that most of the submicron CMOS process design rules do not lend themselves to straightforward linear scaling. The use of lambda-based design rules must therefore be handled with caution in sub-micron geometries. Based on the assumption of: λ half of the minimum feature size 0.75λ worst case misalignment of a mask 1.5λ worst case misalignment mask to mask Gives the following rules for an NFET: 2λ Minimum width of gate 2λ Minimum width of contact λ Minimum enclosure of contact by diff 2λ Minimum extension of poly beyond diff 2λ Minimum space of contact to poly And the following derived rules: 4λ Minimum width of diff 5λ Minimum length of diff In the following, we present a sample set of the lambda-based layout design rules devised for the MOSIS CMOS process and illustrate the implications of these rules on a section a simple layout which includes two transistors

MOSIS Layout Design Rules (sample set) Rule number Description R1 R2 Minimum active area width Minimum active area spacing R3 R4 R5 R6 Minimum poly width Minimum poly spacing Minimum gate extension of poly over active Minimum poly-active edge spacing (poly outside active area) Minimum poly-active edge spacing (poly inside active area) R7 L-Rule 3L 3L 2L 2L 2L 1L 3L R8 R9 Minimum metal width Minimum metal spacing R10 R11 R12 R13 R14 Poly contact size Minimum poly contact spacing Minimum poly contact to poly edge spacing Minimum poly contact to metal edge spacing Minimum poly contact to active edge spacing 2L 2L 1L 1L 3L R15 R16 Active contact size Minimum active contact spacing (on the same active region) Minimum active contact to active edge spacing Minimum active contact to metal edge spacing Minimum active contact to poly edge spacing Minimum active contact spacing (on different active regions) 2L R17 R18 R19 R20 3L 3L 2L 1L 1L 3L 6L

Figure-2.14: Illustration of some of the typical MOSIS layout design rules listed above.

1.9 Design Rules - MOSIS Scalable CMOS (SCMOS) 1.9.1 Introduction 1.9.2 1.9.3 1.9.4 SCMOS Design Rules Well Type SCMOS Options 1.9.1 Introduction MOSIS Scalable CMOS (SCMOS) is a set of logical layers together with their design rules, which provide a nearly process- and metric-independent interface to many CMOS fabrication processes available through MOSIS. The designer works in the abstract SCMOS layers and metric unit ("lambda"). He then specifies which process and feature size he wants the design to be fabricated in. MOSIS maps the SCMOS design onto that process, generating the true logical layers and absolute dimensions required by the process vendor. The designer can often submit exactly the same design, but to a different fabrication process or feature size. MOSIS alone handles the new mapping. By contrast, using a specific vendor's layers and design rules ("vendor rules") will yield a design which is less likely to be directly portable to any other process or feature size. Vendor rules usually need more logical layers than the SCMOS rules, even though both fabricate onto exactly the same process. More layers means more design rules, a higher learning curve for that one process, more interactions to worry about, more complex design support required, and longer layout development times. Porting the design to a new process will be burdensome. SCMOS designers access process-specific features by using MOSIS-provided abstract layers which implement those features. For example, a designer wishing to use secondpoly would use the MOSIS-provided second-poly abstract layer, but must then submit to a process providing for two polysilicon layers. In the same way, designers may access multiple metals, or different types of analog structures such as capacitors and resistors, without having to learn any new set of design rules for the more standard layers such as metal-1. SCMOS is there for portability and simplicity. It is NOT there for fine-tuned layout. Vendor rules may be more appropriate when seeking maximal use of silicon area, more direct control over analog circuit parameters, or for very large production runs, where the added investment in development time and loss of design portability is clearly justified. However the advantages of using SCMOS rules may far outweigh such concerns, and should be considered. 1.9.2 SCMOS Design Rules In the SCMOS rules, circuit geometries are specified in the Mead and Conway's lambda based methodology. The unit of measurement, lambda, can easily be scaled to different fabrication processes as semiconductor technology advances.