You are on page 1of 78

MEMRISTOR DEVICES: FABRICATION, CHARACTERIZATION, SIMULATION,

AND CIRCUIT DESIGN

Thesis

Submitted to

The School of Engineering of the

UNIVERSITY OF DAYTON

In Partial Fulfillment of the Requirements for

The Degree of

Master of Science in Electrical Engineering

By

Chris Yakopcic

Dayton, Ohio

August, 2011
MEMRISTOR DEVICES: FABRICATION, CHARACTERIZATION, SIMULATION,

AND CIRCUIT DESIGN

Name: Yakopcic, Chris

APPROVED BY:

_________________________________ _________________________________
Tarek M. Taha, Ph.D. Guru Subramanyam, Ph.D.
Advisory Committee Chairman Committee Member
Associate Professor Chair and Professor
Electrical and Computer Engineering Electrical and Computer Engineering

_________________________________
Andrew Sarangan, Ph.D.
Committee Member
Associate Professor
Electro-Optics

_________________________________ _________________________________
John G. Weber, Ph.D. Tony E. Saliba, Ph.D.
Associate Dean Dean, School of Engineering
School of Engineering & Wilke Distinguished Professor

ii
©Copyright by

Chris Yakopcic

All Rights Reserved

2011

iii
ABSTRACT

MEMRISTOR DEVICES: FABRICATION, CHARACTERIZATION, SIMULATION,

AND CIRCUIT DESIGN

Name: Yakopcic, Chris


University of Dayton

Advisor: Dr. Tarek M. Taha

Significant interest has been placed on developing systems based on memristors

since the initial fabrication by HP Labs in 2008 [1]. The memristor is a nanoscale device

with dynamic resistance that is able to retain the last programmed resistance value after

power is removed from the device. This property shows that the memristor can be used as

a non-volatile memory component, and has potential to enhance many types of systems,

such as high density memory, and neuromorphic computing architectures.

This thesis presents the fabrication and characterization results obtained based

memristor devices developed at the University of Dayton. In addition, a comparison

between the existing memristor device models was completed to show how the memristor

can be used in a multistate operation. Lastly, circuit designs were completed that

demonstrate the writing and reading of information to and from memristor devices. These

represent the initial steps required in developing electronic systems based on memristors.

A large portion of the work completed in this thesis has been published in [2-4].

iv
ACKNOWLEDGMENTS

My special thanks are in order to Dr. Tarek M. Taha, my advisor, for providing

the time and equipment necessary for the work contained herein, and for directing this

thesis and bringing it to its conclusion with patience and expertise.

I would also like to express my appreciation to everyone who has helped me with

this work. This includes Dr. Andrew Sarangan and Dr. Eunsung Shin, who were

responsible for fabricating the devices characterized in these experiments; Dr. Guru

Subramanyam and Dr. Douglas Hansen, who provided the necessary equipment for

characterizing the devices; and Mark Patterson, who offered his help and expertise when

designing the mask set used for the second device fabrication experiment.

v
TABLE OF CONTENTS

ABSTRACT ....................................................................................................................... iv

ACKNOWLEDGMENTS ...................................................................................................v

LIST OF FIGURES ........................................................................................................... ix

I. INTRODUCTION ............................................................................................................1

II. BACKGROUND .............................................................................................................5

2.1 Memristor Origin ....................................................................................................... 5

2.2 Physical Memristor Discovery .................................................................................. 6

2.2.1 HP Labs Titanium Dioxide Memristor................................................................ 7

2.2.2 Alternative Thin-Film Memristor Designs .......................................................... 9

2.2.3 Spintronic Memristor Devices .......................................................................... 10

2.3 RRAM ..................................................................................................................... 11

2.4 Device Modeling ..................................................................................................... 11

2.5 Circuit Design .......................................................................................................... 12

2.6 Applications ............................................................................................................. 14

III. PHYSICAL DEVICE EXPERIMENT 1 .....................................................................15

3.1 Device Fabrication ................................................................................................... 15

vi
3.2 Experimental Test Setup .......................................................................................... 17

3.3 Experimental Test Results ....................................................................................... 18

3.3.1 Characterization by User Controlled Test Setup ............................................... 18

3.3.2 Characterization by Computer Controlled Test Setup ...................................... 22

3.3.3 Characterization Using Pulse Waveform .......................................................... 25

IV. PHYSICAL DEVICE EXPERIMENT 2 .....................................................................27

4.1 Mask Design for Future Devices ............................................................................. 27

4.1.1 Overall Mask Set ............................................................................................... 28

4.1.2 Device Patterns in Mask Design ....................................................................... 29

4.2 Second Generation Device Fabrication ................................................................... 31

4.3 Second Generation Device Characterization ........................................................... 34

V. COMPARISON OF DEVICE MODELING METHODS ............................................37

5.1 Modeling Equations ................................................................................................. 37

5.2 Equation Based Device Simulation ......................................................................... 40

5.2.1 Device Simulation Based on Linear Model ...................................................... 40

5.2.2 Device Simulation Based on Non-Linear Model 1 ........................................... 42

5.2.3 Device Simulation Based on Non-Linear Model 2 ........................................... 43

5.3 Resistance State Analysis Based on Device Models ............................................... 44

5.3.1 Resistance States in the Linear Device Model .................................................. 44

5.3.2 Resistance States in the Non-Linear Device Models ........................................ 45

vii
5.3.3 Resistance State Comparison ............................................................................ 47

5.4 Device Modeling Based on Fabrication Data .......................................................... 50

VI. CIRCUIT DESIGN BASED ON NON-LINEAR DRIFT MODEL ...........................53

6.1 Circuit for Writing to Memristor ............................................................................. 53

6.1.1 Write Circuit Schematic .................................................................................... 53

6.1.2 Simulation Results for Memristor Write Circuit ............................................... 54

6.2 Circuit for Reading Resistance of a Memristor Device ........................................... 56

6.2.1 Read Circuit Schematic ..................................................................................... 56

6.2.2 Simulation Results for Memristor Read Circuit ................................................ 57

VII. CONCLUSION ..........................................................................................................59

BIBLIOGRAPHY ..............................................................................................................62

viii
LIST OF FIGURES

3.1. Device layout for UDMEM1. .................................................................................... 16

3.2. Device layout for UDMEM2. .................................................................................... 16

3.3. Device layout for UDMEM8. .................................................................................... 16

3.4. Memristors on wafer: (a) alignment of multiple memristors on a wafer, and (b) probe

applied to test a single device. .......................................................................................... 17

3.5. Test results for UDMEM1 with a voltage loop ranging from -4V to 4V. ................. 19

3.6. Test results for UDMEM2 with a voltage loop ranging from -4V to 4V. ................. 19

3.7. I-V curve results for UDMEM2 with voltage ranging from -8V to 5V.................... 20

3.8. Multiple voltage loop test, first pass. ......................................................................... 21

3.9. Multiple voltage loop test, second pass. .................................................................... 21

3.10. Multiple voltage loop test, third pass. ...................................................................... 21

3.11. Test results for UDMEM2 with a voltage from -5 to 5V on the first pass. ............. 23

3.12. Test results for UDMEM2 with a voltage from -5 to 5V on the second pass.......... 23

3.13. Test results for UDMEM2 with a voltage from -5 to 5V on the fifth pass. ............. 23

3.14. Test results for UDMEM8 with a voltage from -4 to 4V on the first pass. ............. 24

3.15. Test results for UDMEM8 with a voltage from -4 to 4V on the second pass.......... 25

3.16. Test results for UDMEM8 with a voltage from -4 to 4V on the first pass. ............. 25

3.17. Test results showing the resistance drop across the memristor as a function of the

number of voltage pulses applied to the device. ............................................................... 26

ix
4.1. Layout of entire mask. ............................................................................................... 28

4.2. Layout for (a) top metal mask, (b) etch mask, and (c) bottom metal mask. .............. 28

4.3. Layout of 4 different areas in the mask set including (a) the larger individual devices,

(b) the crossbar with 100 µm2 overlap, (c) the smaller individual devices, and (d) the

crossbar with 25 µm2 overlap. .......................................................................................... 30

4.4. Span of devices that are included in the larger device set (25 µm2 to 400 µm2). ...... 31

4.5. Span of devices in the smaller devices set (0 µm2 to 35 µm2). .................................. 31

4.6. Device layout for second generation memristor device (contact pads not shown).... 32

4.7. Images of the SiO2 wafer after Pt deposition displaying (a) the contact pad and wire

leading to an isolated device, and (b) the wire passing through the circular definition of a

crossbar device. ................................................................................................................. 32

4.8. Image displaying memristor wafer after deposition of TiO2 and TiO2-x layers. ........ 33

4.9. Images displaying the lithography pattern for the top metal for (a) an isolated device,

and (b) a device in a crossbar. ........................................................................................... 33

4.10. Images of fabricated wafer displaying (a) portion of a memristor crossbar, and (b)

an array of single devices in a wire bonded packaging. ................................................... 33

4.11. Test results for the 200µm2 memristor device. ........................................................ 36

4.12. Test results for the 100µm2 crossbar memristor device. .......................................... 36

5.1. Simulation results for testing liner drift velocity model. The top left plot shows the

current and voltage waveforms with respect to time, and the top right plot shows the I-V

characteristic. The bottom left plot shows position of the barrier between the titanium

oxide layers and the bottom right plot shows resistance as a function of time. Simulation

x
parameters are, RON=100Ω, ROFF=1000Ω, D=46nm, µD=10-14m2V-1s-1, w0/D=.2, with a

triangle voltage input where f=2Hz and A=.25V. ............................................................. 41

5.2. Simulation results for testing the Joglekar non-linear drift velocity model with p=1.

The top left plot shows the current and voltage waveforms w. r. t. time, and the top right

plot shows the I-V characteristic. The bottom left plot shows position of the barrier

between the titanium oxide layers and the bottom right plot shows resistance as a function

of time. Simulation parameters are, RON=100Ω, ROFF=1000Ω, D=37nm, µD=10-14m2V-1s-


1
, w0/D=.2, with a triangle voltage input where f=2Hz and A=.25V. ................................ 42

5.3. Simulation results for testing the Biolek non-linear drift velocity model with p=1.

The top left plot shows the current and voltage waveforms w. r. t. time, and the top right

plot shows the I-V characteristic. The bottom left plot shows position of the barrier

between the titanium oxide layers and the bottom right plot shows resistance as a function

of time. Simulation parameters are, RON=100Ω, ROFF=1000Ω, D=32nm, µD=10-14m2V-1s-


1
,w0/D=.2, with a triangle voltage input where f=2Hz and A=.25V. ................................. 43

5.4. Plot on the left shows number of resistance levels achieved using the linear drift

model for D=90-115nm where RON/ROFF=10 for current pulse differentials given as 5, 10,

and 15%. Plot on the right shows same data for RON/ROFF=50 and D=42-55nm. Other

simulation parameters are, RON=100Ω, µD=10-14m2V-1s-1, w0/D=.2, with a triangle voltage

input where f=4Hz and A=2V. .......................................................................................... 45

5.5. Plot on the left shows number of resistance levels achieved using the Joglekar non-

linear drift model for D=60-120nm where RON/ROFF=10 for current pulse differentials

given as 5, 10, and 15%. Plot on the right shows same data for RON/ROFF=50 and D=30-

xi
50nm. Other simulation parameters are, RON=100Ω, µD=10-14m2V-1s-1, w0/D=.2, with a

triangle voltage input where f=4Hz and A=2V. ................................................................ 46

5.6. Simulation results for testing the Joglekar non-linear drift velocity model with p=1.

The top left plot shows the current and voltage waveforms w. r. t. time, and the top right

plot shows the I-V characteristic. The bottom left plot shows position of the barrier

between the titanium oxide layers and the bottom right plot shows resistance as a function

of time. Due to the strong memristive effect, device operates as a linear resistor after 1.5s.

Simulation parameters are, RON=100Ω, ROFF=1000Ω, D=50nm, µD=10-14m2V-1s-


1
,w0/D=.2, with a triangle voltage input where f=4Hz and A=2V. .................................... 47

5.7. Plot on the left shows number of resistance levels achieved using the Biolek non-

linear drift model for D=50-110nm where RON/ROFF=10 for current pulse differentials

given as 5, 10, and 15%. Plot on the right shows same data for RON/ROFF=50 and D=30-

50nm. Other simulation parameters are, RON=100Ω, µD=10-14m2V-1s-1, w0/D=.2, with a

triangle voltage input where f=4Hz and A=2V. ................................................................ 47

5.8. Comparison of results for resistance level data presented in previous section using a

minimum pulse differential of 5%. ................................................................................... 49

5.9. Device conductance as a function of doped layer thickness when using simulations

based on equation (1). ....................................................................................................... 49

5.10. Simulation results for numerically based model that incorporates programming

threshold displaying (a) the voltage input, (b) the memristor state, and (c) the dynamic

resistance value. ................................................................................................................ 52

6.1 MOSFET based memristor write circuit. .................................................................... 54

xii
6.2. Memristor response to write circuit simulation with voltage applied from V1. Top

left plot shows the voltage across the memristor, when the 5V pulses are applied, the top

right plot shows the current in the memristor and the voltage pulse are applied, and the

bottom plot shows the shift if the state variable w(t) relative to the thickness of the

device. ............................................................................................................................... 55

6.3. Memristor response to write circuit simulation with voltage applied from V2. Top

left plot shows the voltage across the memristor, when the 5V pulses are applied, the top

right plot shows the current in the memristor and the voltage pulse are applied, and the

bottom plot shows the shift if the state variable w(t) relative to the thickness of the

device. ............................................................................................................................... 55

6.4 Comparator based read circuit for memristor. ............................................................ 57

6.5. Read circuit simulation. The top plot shows the input voltage, the middle plot shows

the voltage generated between the memristor and the 75k resistor, and the bottom plot

shows the output of each of the comparators. ................................................................... 58

xiii
CHAPTER I

INTRODUCTION

For nearly 180 years, it has been accepted that there are three fundamental

passive circuit elements, the resistor (1827), the capacitor (1745), and the inductor

(1831). In 1971 Dr. Leon Chua theorized that mathematically there should be a fourth

fundamental circuit element based on the symmetry of the equations that govern passive

circuit theory [5]. Dr. Chua called this device the memristor (short for memory-resistor),

and in 2008 Hewlett Packard Labs published results of the device in physical form [1].

The physical memristor is nanoscale device that has unique properties that can be

used to greatly improve existing electronic systems and computing architectures. The

memristor can be thought of as a time varying resistor where the resistance changes due

to the summation of current that has passed through the device [5]. When the current

flowing through the device is zero, the summation of current becomes constant, and thus

the resistance remains unchanged. This shows that the memristor can be used as a non-

volatile memory component.

Furthermore, the dynamics of a memristor closely resemble those of a synapse in

brain tissue. Just as the values of synaptic weights change with the application of neural

spikes, the resistance value of a memristor can be changed with the application of a

voltage pulse. This could provide significant advancements in the field of neuromorphic

1
computing as electronic systems using memristors could be fabricated with a device

density similar to that of the human brain [6].

Since the memristor‘s physical discovery, several institutions have published

memristor device fabrications using a variety of different materials and device structures

[2, 7-13]. Different device structures are still being developed to determine which

memristor device would be the best option for commercial use. This is based on many

factors such as size, switching speed, power consumption, switching longevity, and

CMOS compatibility.

Several device models have also been proposed that can be used simulate

memristor based circuits [1, 14, and 15]. The circuits published thus far involve methods

of writing to and reading from memristors [16-19], and using a memristor based circuit as

a component in a neuromorphic system [20, 21]. The memristor devices, memristor

models, and basic circuit designs will serve as the first steps in developing electronic

systems and computing architectures based on memristors.

The work presented in this thesis involves memristor device fabrication and

characterization, memristor model comparisons, and read and write circuit designs. Two

different memristor fabrication experiments are described. The first was based on a set of

wafers containing an array of isolated devices with a large cross-sectional area (375 by

375µm) where each wafer was fabricated with different oxide thicknesses. The results of

this experiment showed that the wafer with a total oxide thickness of 100 nm provided

the most reliable devices. A large hysteresis pattern could be seen when the device was

positively biased. The work based on this experiment was published in [3, 4]. The second

memristor fabrication experiment was based on a much more complex wafer. A mask set

2
was designed to produce isolated memristor devices with variable cross-sectional area

(2.5µm2 to 400µm2) in addition to small crossbar arrays containing 36 devices at a cross-

sectional area of ether 25µm2 or 100µm2. The results of the wafer based on this design

provided a stronger hysteresis pattern in both the positive and negative voltage regimes.

A comparison of three existing memristor device models was performed to show

how the memristor could be used as a multistate memory device for use in a high density

memory system or a synaptic weight in a neural system. The three models compared

were published in [1, 14, and 15] and were developed on the basis that a memristor acts

as two variable resistors in series where the dynamic resistance component is determined

by a state variable. The state variable assigns a numerical value to the physical changes in

the device that relate to a change in resistance. The motion of the state variable differs in

each of these models. The simulations completed in this experiment examine how many

distinguishable resistance values can be observed after a series of voltage pulses are

applied to the device model. The results showed that the model in [15] provided the

maximum amount of observable resistance levels, although the models have a large

amount of discrepancy when comparing to actual characterization data. The work

completed based on the model comparisons was published in [2]. An alternative

modeling approach was developed based on characterization data provided in [10], and

was developed using a lookup table approach.

Circuits were then developed to write information and to read information from

memristors. Simulations of the circuit designs were performed in SPICE using the

memristor model published in [15]. The simulation results for the write circuit show that

the state (conductivity) of a memristor can be increased or decreased when a positive

3
voltage pulse is applied to one of two inputs. The read circuit shows how a system of

comparators can be used to quantize the resistance value of a memristor into a discrete

number of states for digital processing.

Chapter 2 provides a background section that describes the work completed based

on memristors since it was first theorized by Dr. Chua [5]. Chapter 3 presents the

fabrication techniques and results for the first memristor fabrication experiment carried

out at the University of Dayton. Chapter 4 describes the mask design for the more

complex set of memristors along with fabrication information and test results. Chapter 5

shows how existing memristor models and characterization data may be used to predict

the behavior of a memristor when used as a multi-state device. Chapter 6 presents circuit

designs and simulation results for reading and writing to memristor devices. Lastly,

Chapter 7 provides a conclusion of the results.

4
CHAPTER II

BACKGROUND

2.1 Memristor Origin

The memristor was first theorized by Dr. Leon Chua in 1971 [5] as the fourth

passive fundamental circuit element. Based on the symmetry of the equations that govern

the resistor, capacitor and inductor, Dr. Chua hypothesized that fourth device should exist

that holds a relationship between magnetic flux and charge. This would complete the

circle where the resistor holds the relation between current and voltage, the inductor

holds the relationship between current and flux, and the capacitor holds the relationship

between voltage and charge. Dr. Chua proved that the abilities of the memristor could not

be duplicated by any of the other three passive elements, and that an active circuit that

mimics the functionality of the theorized device would require approximately 25

transistors. The memristor was predicted to be a device with a dynamic resistance that is

determined by the integral of current flowing through it. Since this is an integral

relationship, applying zero current would result in a constant charge, thus leaving the

resistance constant. This shows that the theorized memristor possesses the ability to

retain a resistance value even after the power source is removed from the device. Unlike

the capacitor and inductor, this is not an energy storage device, so the voltage must equal

zero whenever the current goes to zero. This causes the I-V curve of the device to

5
produce a pinched hysteresis loop, revealing that the device has a memory effect

associated with it.

A few years later, Dr. Chua and Dr. Sung Mo Kang, produced further research on

the theoretical memristor device, claiming the memristor to be a subset of a broader class

defined as memristive systems [22]. Like memristors, all memristive systems must have a

zero output whenever the input is zero, and this is what distinguishes them from arbitrary

dynamical systems. Chua and Kang found publications dating back to the early 1900s

that described systems that could be classified as memristive, but this connection had

never been realized. The memristor was considered nothing more than a mathematical

curiosity until the first successful fabrication of a memristor device was published in

2008 [1].

2.2 Physical Memristor Discovery

The first fabrication of a memristor device was developed by a research team at

HP Labs lead Dr. Stanley Williams [1]. The research team at HP Labs was formed in

1995, and they initially started working on ways to further increase computing power

once transistors reach their minimum size constraint. They decided the solution should be

a nanoscale crossbar series of switches, which would be designed so the system could

still function even if some of the components were defective [23].

The initial device structure for the switches at each crossbar junction was a layer

of platinum dioxide, followed by a monolayer film of switching molecules, and finally a

layer of titanium. When considering the platinum crossbars as electrodes, the resulting

device was defined within this 5 layer structure. As the internal mechanism in the device

6
was studied, it became apparent that the oxygen atoms in the device migrated through the

monolayer, and formed titanium dioxide on the other side. They also found that the

titanium dioxide was formed in two different patterns, a stoichiometric layer, and a

slightly oxygen deficient layer. The monolayer was then removed and a fabrication

method was determined that would yield reliable devices that contained both the

stoichiometric and oxygen deficient layers of titanium oxide [23]. Testing of these

devices showed that the resistance of the device did not change when the voltage source

was removed, and although it wasn‘t circular, the device provided an I-V characteristic in

the form of a pinched hysteresis loop [1]. Following this discovery, HP Labs has

produced extensive research based on the fabrication, modeling, and dynamics of

titanium dioxide memristor devices [24-32].

2.2.1 HP Labs Titanium Dioxide Memristor

The structure of the thin-film titanium dioxide memristor was decided to be a

stoichiometric layer of titanium dioxide (TiO2) and an oxygen deficient layer (TiO2-x)

sandwiched between two platinum electrodes [1]. The device was determined to be

functioning by the migration of the charged vacancies in the oxygen deficient layer.

When a positive voltage was applied to the device, the oxygen vacancies expanded. This

created a thicker oxygen deficient layer, thus reducing the resistance of the device.

Conversely, when a negative voltage was applied to the device, the oxygen vacancies

contracted and the resistance of the device increased. HP Labs created a simple

mathematical model that described this phenomenon based on two dynamic resistors and

a differential equation that modeled the migration of oxygen vacancies [1].

7
HP Labs has since published many more papers based on the characterization of

the TiO2 memristors that they developed [24-32]. One of these papers describes a more

complex test setup for characterizing memristor devices [26]. A series of pulses was

applied to the device with a constant voltage, but a variable pulse width. This was done to

characterize the state variable that determined the resistance of the device. The state

variable was then fit to the Simmon‘s tunneling equation for a metal-insulator-metal

(MIM) junction with image forces [33]. The mechanism for resistance change was then

described as a tunnel barrier with variable thickness.

Further work was done on this device to see if Joule heating had an impact on the

resistive switching process of a memristor device [27]. These results also concluded that

the resistance change is determined by the modulation of the width of a tunnel barrier.

Furthermore, as temperature significantly changed the resistance value, temperature can

potentially be seen as a hidden state variable within the device.

Dr. Paul Strachan et al. published a paper that describes the electroforming

process in the memristive devices [30]. These devices were fabricated using TiO2

sandwiched between two platinum electrodes. The application of a high voltage changed

the makeup of the device, leaving a bubble of oxygen deficient material in the thin-film.

The TiO2-x channel formed only in a small portion of the device, and a small layer of

TiO2 was left between the oxygen deficient channel and the platinum conductor which

supports the possibility of a tunnel gap. Dr. Strachan et al. described the oxygen deficient

layer as possibly having a crystal structure of Ti4O7, [32]. These results were supported

by a group led by Dr. Deok-Hwang Kwon at Seoul National University as they found the

same crystal structure present in their devices [34].

8
2.2.2 Alternative Thin-Film Memristor Designs

Since HP labs discovered the memristor, several academic institutions, including

the University of Dayton, have published results of memristor devices. These devices

show that the memristor can be developed using a wide variety of materials and device

structures. The reasons for the memristive effect include, but are not limited to, the

motion of oxygen vacancies [1], the motion of silver dopants [7], and the state change in

chalcogenide materials [9]. A complete study of which materials contain the required

properties to be considered memristive has not yet been published. Until then, the

reasoning for choosing each material in the devices appears to be on a case by case basis.

Sung Hyun Jo and Wei Lu at the University of Michigan fabricated memristors

published in [7] that are based on an amorphous silicon and silver. As voltage is applied,

Ag ions are pulled into the a-Si layer to lower the resistance of the device. The thickness

of the a-Si was 120nm and the cross-sectional area of the devices was 100 by 100nm.

Sung Hyun Jo et al. also fabricated devices using a slightly different method in

[8]. In this paper, a-Si is co-sputtered with Ag at a ratio that changes as the layer is

formed on the wafer. This produces a film that has a low concentration of Ag at one end

and a higher concentration of Ag at the other. These devices have an area of 100 by

100nm. The devices had significantly lower switching times of about 5-10ns when

compared to the device previously described in [7].

Antonio S. Oblea et al. has fabricated chalcogenide based devices where Ag2Se

and Ag were placed between tungsten electrodes, [10]. Ge2Se3 was placed between each

layer to help the tungsten adhere, and to make it possible to deposit the Ag layer. The

devices had a diameter of 180nm and were fabricated by first depositing a substrate, then

9
etching vias where the devices would be deposited. The electrodes were formed with a

much larger area of 80 by 80µm.

Nadine Gergel-Hackett et al. developed a flexible memristor that was deposited

on a laserjet transparency [11]. The device is formed by letting a titanium isopropoxide

solution hydrolyze to form a 60nm layer of amorphous TiO2 placed between Al

electrodes. The device is large with cross-sectional area of 4mm2. The device shows

on/off rations greater that 10,000:1 and has data retention capabilities of about 14 days.

This suggests that the memristive effect can be seen in devices with sizes that exceed the

nanometer scale.

A group led by T. Prodromakis conducted a study that tested memristance on

devices that range from 1 by 1µm2 to 200 by 200µm2 to further show that the memristive

effect is not only present in the nanoscale. The results show that hysteresis size and the

ratio between the maximum and minimum resistance increase as the device area

decreases [12].

Titanium dioxide memristors have also been fabricated using a slightly different

approach; using a TiO2+x layer in addition to the stoichiometric one. In this case, the

abundance of oxygen atoms is negatively charged which theoretically creates a memristor

with charges that move opposite to the TiO2-x structure. T. Prodromakis et al. developed

TiO2/TiO2+x devices that exhibit memristive behavior with an active area of 1 to 25µm2

[13].

2.2.3 Spintronic Memristor Devices

In addition to the development of thin-film memristors, attempts have been made

to develop spintronic memristors [35-37]. The resistance is variable dependent on the

10
magnetoresistance and the spin torque induced magnetized motion. The device was

designed using two ferromagnetic layers, one reference layer and one free. The free layer

is split into two sections magnetized in opposite directions. The position of the barrier

dividing the layers of opposite magnetization directions is what determines the resistance

of the device. The advantage of this system is that it can be integrated into CMOS just as

magnetic RAM that is currently commercialized.

2.3 RRAM

RRAM is a device that works similarly to memristor devices, as it is a form of

non-volatile memory that works by changing the resistance of a dielectric material. They

are primarily made using an oxide layer between two electrodes. A common dielectric

used in these devices is hafnium oxide where one example shows a device containing a

5nm HfO2 layer [12]. The RRAM devices that have been fabricated generally have high

and low resistance states that represent logic 0 or 1, as opposed to memristors that could

potentially hold many resistance states.

2.4 Device Modeling

In addition to physical device development and characterization, memristor

devices have also been modeled based on the equations first proposed by HP Labs [1, 14,

and 15]. The memristor device was first modeled by two variable resistors in series. The

resistance of each was determined by the instantaneous thickness of the two thin-film

layers made up of high resistivity TiO2 and low resistivity TiO2-x. The thickness of the

TiO2-x layer is modeled using a state variable that changes based on the net charge

11
applied to the device. Some additional analysis was done in [14, 15] that expands on the

equations presented by HP Labs in [1].

HP Labs released a memristor model in [1] that describes the charges in the doped

layer to move with linear dopant drift. Since then, Joglekar and Wolf expanded on the

model in [14] by defining a non-linear drift velocity. Furthermore, Zdenek Biolek et al.

describe an alternative non-linear drift model for the memristor [15]. These models do

not take into account the non-linear tunneling and heating effects described in [26, 27].

Further modeling has been completed based on the chalcogenide device

developed at Boise State University [9, 39]. These models were developed so that the I-V

characteristics of both the fabricated device and the model match very closely. The

drawback of this type of model is that it conforms to the device data for a limited number

of voltage inputs.

2.5 Circuit Design

Circuit designs have been published that show how the memristor could be added

to passive circuit theory using RLCM designs [14, 40]. This is an addition to RLC

passive circuit theory that involves a non-linear, time-varying element. First and second

order systems were developed as a memristor based approach to design chaotic

oscillators.

Circuits have also been published that use the memristor as a programmable

resistor in an analog circuit [41]. Since high frequency signals with zero net flux do not

disrupt the state of a memristor, a DC signal is applied to set the resistance value, and

then the AC signal propagates with stable memristor operation. This technique has been

12
applied to programmable gain amplifiers, where the memristor is in place of the feedback

resistor.

Dr. Eshraghian et al. proposed a circuit design for a content addressable memory

(CAM) cell that utilizes memristors [42]. For this design, the memristor is combined with

CMOS transistors to provide a CAM cell with a 45% reduction in area when compared

with an SRAM equivalent. This new MCAM cell also provides a 74% reduction in power

consumption, although the write operation time is substantially longer due to the slow

switching time of the memristor.

Circuits have also been proposed that use memristors in cellular neural network

(CNN) structures [20, 21]. The memristors act as synaptic weights, where the resistance

represents signal strength. These circuits are based on a memristor/CMOS hybrid design

that would replace a transistor based design.

A large number of read and write circuits were developed to control the information

processing aspect of memristor devices. Some of these circuits involve using op amps as

analog comparators that determine the state of a memristor [16-18]. To change the state

of the memristor, a voltage pulse is applied where a transistor switching mechanism is

used to control whether to read or write to the device. The read and write circuits

proposed in [19] are complex designs that have the ability to modify the memristor

resistance to a pre-defined set of 8 different values.

13
2.6 Applications

The three most promising applications of the memristor are non-volatile memory

[8, 16-19, 43], neuromorphic architectures [6, 7, 20, 21, 44, 45], and implication logic

used for digital circuit design [46].

Non-volatile memory can be designed based on the dynamic resistance of the

memristor device. For a simple one bit storage application, a high resistance could

represent logic 0, and a low resistance could represent logic 1. Memristor based non-

volatile memory could potentially provide much more data storage per footprint area

when compared the current technology [23]. Higher density can be achieved using the

memristor as a multistate memory device [19].

Memristor based neuromorphic systems could be developed where the memristor

holds a value representing the synaptic weight in a neural network [6]. Several papers

have shown memristor fabrications where the memristor is capable of storing several

different resistance values based on the application of a string of voltage pulses [6, 7, and

10].

Due to the dynamics of the memristor, logic design based on these devices is best

done using material implication. A set containing the implication operator and the not

operator, is computationally complete, as is the set {and, not}. The change in the logic

operators is required as this would create a mathematical system that more closely

resembles how memristors work [32].

14
CHAPTER III

PHYSICAL DEVICE EXPERIMENT 1

The following chapter discusses the fabrication and characterization of TiO2

memristor devices. Of the all of the fabricated devices, the three most successful

fabrications are known as UDMEM1, UDMEM2, and UDMEM8. The fabrication

method of these devices is discussed, as well as the experimental setup and results

obtained when testing the devices.

3.1 Device Fabrication

UDMEM1, UDMEM2 and UDMEM8, differ in the thickness of the dielectric

layers. UDMEM1 includes a 900 nm titanium oxide layer and a 440 nm oxygen deficient

titanium oxide layer sandwiched between top and bottom electrodes (see Figures 3.1, 3.2,

and 3.3). UDMEM2 has the same device structure except that the titanium dioxide layer

was reduced to 450 nm and the TiO2-x layer was reduced to 220 nm. UDMEM8 is a

similar device with thinner titanium oxide layers (50 nm for titanium oxide and 50 nm for

oxygen deficient layers). In each case, the bottom layer on the device is a 400 nm layer of

Si with resistivity of 1 to 10 Ωcm.

The titanium oxide layers were formed by magnetron sputtering at an oxygen

pressure of 14 mTorr. The UDMEM1, UDMEM2, and UDMEM3 devices were formed

at a temperature of 200°C, 350°C, and 400°C respectively.

15
For each of the devices, the bottom electrode was a typical metal stack used for

Ohmic contact on low resistivity Si – about 500 nm thick (Ti/Al/Ni/Au) capped with a

100 nm Pt layer. The top layer was a Ti/Au layer with a thickness of about 800 nm. The

bottom metal layer was rapid thermal annealed at 850°C for 30 seconds to create the

Ohmic contact to the Si substrate. Three wafers, one for each of the devices (UDMEM1,

UDMEM2 and UDMEM8), were fabricated with several devices on each wafer.

Metal 2 800 nm
TiO2-x 440 nm
900 nm
TiO2
Metal 1 500 nm

Low Resistivity Si 400 nm

Figure 3.1. Device layout for UDMEM1.

Metal 2 800 nm
TiO2-x 220 nm
450 nm
TiO2
Metal 1 500 nm

Low Resistivity Si 400 nm

Figure 3.2. Device layout for UDMEM2.

Metal 2 800 nm

TiO2-x 50 nm
TiO2 50 nm

Metal 1 500 nm

Low Resistivity Si 400 nm

Figure 3.3. Device layout for UDMEM8.

Figure 3.4 (a) shows how multiple devices are aligned on the UDMEM2 wafer.

Figure 3.4 (b) shows a single device being examined using a probe under a microscope.

To test each device, a probe was applied to the top electrode, the 800 nm Ti/Au (see Figs.

16
1 and 2). The device was grounded through the low resistivity Si, which was connected to

the bottom layer of the device.

(a) (b)

Figure 3.4. Memristors on wafer: (a) alignment of multiple memristors on a wafer, and (b) probe applied to
test a single device.

3.2 Experimental Test Setup

Two different methods were used to obtain the I-V curves for the devices.

UDMEM1 and UDMEM2 were tested using a Keithley 2400 SourceMeter, so that a

voltage could be applied and the current read to generate device I-V curves. Each voltage

was entered in step intervals by hand, first increasing to the maximum voltage, then

decreasing past zero to the minimum voltage, and back to zero. This provides the ability

to use a DC voltage source to apply voltage in a cyclic fashion.

The UDMEM2 device was again tested along with UDMEM8 using a Princeton

Applied Research computer controlled parameter analyzer in a cyclic voltammetry setup.

UDMEM1 was not subject to further testing as it had inferior performance when

compared to UDMEM2 in the initial testing. This method allowed many more points to

be collected in a more accurate and efficient manner. The current for a number of voltage

steps was recorded to obtain the resulting I-V plots.

17
3.3 Experimental Test Results

The tests conducted on the fabricated memristors were done to first determine that

the devices are acting as memristors, then to determine the extent to which these devices

can successfully act as electrical synapses. First, tests were conducted to produce an I-V

curve, both by hand and using a parameter analyzer. Then, repetitive pulses were applied

to determine if this can effectively change the resistance of the device.

3.3.1 Characterization by User Controlled Test Setup

The current-voltage results for UDMEM1 and UDMEM2 done using the Keithley

SourceMeter are now examined to determine if the memristor device functions properly.

First, a single device on each of the wafers was selected and a voltage loop was applied.

In this test, a voltage loop is considered to be from 0V to 4V, then 4V to -4V, and back to

0V. Figure 3.5 shows the I-V curve for the UDMEM1 sample. A slight hysteresis loop

can be seen where positive voltage was applied to the device. The current flowing

through the device was slightly higher on the voltage decrease as opposed to the initial

increase. Figure 3.6 shows the I-V results for the UDMEM2 sample. This sample had a

thinner titanium dioxide layer, and had a larger hysteresis loop when comparing to the

UDMEM1 sample. A small hysteresis can also be seen where the voltage is negative. The

conductivity of UDMEM2 was slightly higher due to a thinner oxide layer.

18
Test 1: I-V UDMEM1 - .2V step -4:4V, SubTH=400um, A = .0014cm 2
30
Voltage Increase
25 Voltage Decrease

20

Current (  A)
15

10

-5
-4 -2 0 2 4
Voltage (V)
Figure 3.5. Test results for UDMEM1 with a voltage loop ranging from -4V to 4V.

Test 2: I-V UDMEM2 - .2V step -4:4V, SubTH=200um, A = .0014cm2


70
Voltage Increase
60 Voltage Decrease

50

40
Current (  A)

30

20

10

-10
-4 -2 0 2 4
Voltage (V)
Figure 3.6. Test results for UDMEM2 with a voltage loop ranging from -4V to 4V.

The UDMEM2 sample was then tested with an increased voltage range to see the

changes in the hysteresis loop. The voltage loop was set from 0V to 5V, 5V to -8V, then -

8V to 0V. The upper limit on the voltage in this test was 5V because the meter used to

determine current can only measure up to 105µA. At 5V, the current measurement was at

a maximum so the voltage was then decreased. When the voltage was decreased, the

current through the device peaked at -10.51µA. These results are plotted in Figure 3.7. It

can be seen that the size of the hysteresis loop increased as voltage range tested became

greater.

19
Test 3: I-V UDMEM2 - .5V step -8:5V, SubTH=200um, A = .0014cm2
120
Voltage Increase
100 Voltage Decrease

80

Current (  A)
60

40

20

-20
-8 -6 -4 -2 0 2 4 6
Voltage (V)
Figure 3.7. I-V curve results for UDMEM2 with voltage ranging from -8V to 5V.

The next test was designed to see how the hysteresis loop changed as multiple

voltage loops were applied in succession. The UDMEM2 sample was selected for this

test, as it provided the superior results when comparing the data in the previous plots. The

voltage applied was increased from 0V to 5V, from 5V to -12V, then -12V to 0V. Since

these tests were done on a different device on the UDMEM2 wafer, the results look

different when comparing to the UDMEM2 test in Figure 3.7, especially when the

voltage was negative. This device had a much higher current flow in the negative region

as well as a larger hysteresis loop. This suggests that the wafer is not uniform, and

different devices behave differently in different locations throughout the wafer. The

following three plots show the I-V results when three consecutive voltage loops were

applied to the device. Figure 3.8 shows the I-V curve for the first voltage loop, Figure 3.9

shows the second, and Figure 3.10 shows the third.

20
Test 4: I-V UDMEM2 - .5V step -12:5.5V, A = .0014cm2 First Pass
120
Voltage Increase
100
Voltage Decrease
80

60

Current (  A)
40

20
0
-20

-40

-60

-80
-10 -5 0 5
Voltage (V)
Figure 3.8. Multiple voltage loop test, first pass.

Test 4: I-V UDMEM2 - .5V step -12:5.5V, A = .0014cm2 Second Pass


120
Voltage Increase
100
Voltage Decrease
80

60
Current (  A)

40

20
0
-20

-40

-60

-80
-10 -5 0 5
Voltage (V)
Figure 3.9. Multiple voltage loop test, second pass.

Test 4: I-V UDMEM2 - .5V step -12:5.5V, A = .0014cm2 Third Pass


120
Voltage Increase
100
Voltage Decrease
80

60
Current (  A)

40

20
0
-20

-40

-60

-80
-10 -5 0 5
Voltage (V)
Figure 3.10. Multiple voltage loop test, third pass.

21
When comparing Figures 3.8, 3.9, and 3.10, it can be seen that the hysteresis loop

became less uniform as the second and third voltage loops were applied. The three I-V

curves appeared to be most predictable when the voltage was positive. When the voltage

was negative, the hysteresis changed much more between each pass and degraded more

severely. This test clearly shows the hysteresis loop with an intersection when the current

and voltage are both 0. This shows that the necessary conditions are met for this device to

be considered memristive.

3.3.2 Characterization by Computer Controlled Test Setup

After testing the devices using the Keithley setup, the computer controlled

parameter analyzer was used to obtain further I-V results. First, a single device on the

UDMEM2 wafer was selected and a voltage loop was applied similar to when using the

Keithley SourceMeter. In this test, a voltage loop was considered to be from 0V to 5V,

then 5V to -5V, and back to 0V. This loop was applied to the device five times and each

pass was observed to see how the I-V curves changed. Figure 3.11 shows the I-V curve

for the first pass on the UDMEM2 sample. A slight hysteresis loop can be seen when a

positive voltage is applied to the device. Figure 3.12 shows the I-V curve for the second

pass when testing the UDMEM2 sample. These results show that the size of the

hysteresis loop increased when compared to the first pass. Then, the hysteresis started to

degrade after the second pass, and the resulting I-V curve after the fifth pass can be seen

in Figure 3.13. In all three cases, the device was only conductive when positively biased.

22
I-V Curve for UDMEM2 Pass 1
10

Current (  A)
4

-2
-5 -4 -3 -2 -1 0 1 2 3 4 5
Voltage (V)

Figure 3.11. Test results for UDMEM2 with a voltage from -5 to 5V on the first pass.

I-V Curve for UDMEM2 Pass 2


18

16

14

12

10
Current (  A)

-2
-5 -4 -3 -2 -1 0 1 2 3 4 5
Voltage (V)

Figure 3.12. Test results for UDMEM2 with a voltage from -5 to 5V on the second pass.

I-V Curve for UDMEM2 Pass 5


10

6
Current (  A)

-2
-5 -4 -3 -2 -1 0 1 2 3 4 5
Voltage (V)

Figure 3.13. Test results for UDMEM2 with a voltage from -5 to 5V on the fifth pass.

23
The test results for the UDMEM8 sample were similar to UDMEM2, except the

hysteresis loop was slightly larger. Also, the device was more conductive since the

titanium dioxide layers in this device were thinner than in the UDMEM2 sample. The

initial increase in conductivity happened much faster in this sample when looking at the

data for the first voltage loop pass in Figure 3.14. Figure 3.15 shows how the size of the

hysteresis loop increased during the second voltage pass. The fifth pass in Figure 3.16

shows degradation in the hysteresis much like in the UDMEM2 device. The UDMEM8

sample also shows that conductivity was only present when the device was positively

biased. These tests both show the hysteresis loop with zero current at a zero voltage bias.

I-V Curve for UDMEM8 Pass 1


100

80

60
Current (  A)

40

20

-20
-2 -1 0 1 2 3 4
Voltage (V)

Figure 3.14. Test results for UDMEM8 with a voltage from -4 to 4V on the first pass.

24
I-V Curve for UDMEM8 Pass 2
120

100

80

Current (  A)
60

40

20

-20
-2 -1 0 1 2 3 4
Voltage (V)

Figure 3.15. Test results for UDMEM8 with a voltage from -4 to 4V on the second pass.
I-V Curve for UDMEM8 Pass 5
120

100

80
Current (  A)

60

40

20

-20
-2 -1 0 1 2 3 4
Voltage (V)

Figure 3.16. Test results for UDMEM8 with a voltage from -4 to 4V on the first pass.

3.3.3 Characterization Using Pulse Waveform

The next test shows the UDMEM2 device‘s reaction to voltage pulses as opposed

to a DC bias. Figure 3.17 shows the results when applying pulses to the device. This test

was done by applying a 2% duty cycle pulse oscillating between 0 and 4V to the

memristor, with a frequency in the range of 3-50Hz. The number of pulses was

determined by multiplying the frequency of the pulse waveform by the amount of time it

was applied to the device. The source was applied for a set amount of time and then an

Ohmmeter was used to measure the resistance across the device as the signal was turned

25
off. This resistance was then compared to the resistance of the device before the source

was applied to obtain the total resistance drop. This data supports the ultimate goal of

having a memristor device model brain function, as the resistance was reduced with the

application of pulse waveforms.

Resistance Drop as a Function of the Number of Voltage Pulses Applied


0

-10
Change in Resistance (k  )

-20

-30

-40

-50

-60
0 200 400 600 800 1000
Number of Pulses Applied (t*f)
Figure 3.17. Test results showing the resistance drop across the memristor as a function of the number of
voltage pulses applied to the device.

26
CHAPTER IV

PHYSICAL DEVICE EXPERIMENT 2

A mask set has been developed to fabricate a more complex set of memristor

devices. This mask set makes it possible to fabricate much smaller devices in terms of

footprint area, as well as memristor crossbar arrays. The fabrication method for the

second generation of devices is discussed, as well as the experimental test results for the

first wafer fabrication using this new method.

4.1 Mask Design for Future Devices

The second generation of memristor devices is based on a new mask set that

makes it possible to fabricate devices that are isolated, as well as devices arranged in a

crossbar pattern. The layout of the entire mask can be seen in Figure 4.1. The device area

when using this mask is determined by the overlap between the top and bottom metal

layers, and the overlap area ranges from 0µm2 to 400 µm2 for different the device

patterns.

27
Figure 4.1. Layout of entire mask.

4.1.1 Overall Mask Set

The mask set contains three different masks, the top metal, the bottom metal, and

an etch mask. Figure 4.2 shows each mask it the mask set. It can be seen that all of the

probing pads are present on the top metal mask, and only the pads connecting to the

bottom electrode on the device are present on the bottom metal mask. The etch mask is

used to provide access to the bottom metal, as vias will be etched wherever the pads

connected to the bottom electrode need to be accessed from the top of the device.

(a) (b) (c)


Figure 4.2. Layout for (a) top metal mask, (b) etch mask, and (c) bottom metal mask.

28
4.1.2 Device Patterns in Mask Design

Each mask is divided in to 4 different sections with an area of 1 cm2, and 8 of

each section are spread out over the entire mask. The four different sections of the mask

are displayed in Figure 4.3. There are two different crossbar designs and two different

sections of individual devices. Figure 4.3 (a) contains 9 different individual devices that

range from 25µm2 (rightmost device) to 400 µm2 (leftmost device). Each of these device

patterns resides between two larger patterns that will act as pads for the device to be

probed. It can be seen that there are four rows of device patterns where each column

represents a different device size. The pad above the device acts as the top electrode, and

the pad below the device provides access to the bottom electrode. The actual memristor

device is defined as a circular overlap between the top and bottom electrodes. This layout

can be seen in more detail in Figure 4.4. The layout in Figure 4.3 (c) is a similar pattern,

except that the devices are smaller in size and are defined by the overlap of two wires

with a width of 5 µm as opposed to a circular overlap. The device size ranges from 2.5

µm2 to 35 µm2 with an additional device with zero overlap area. This was done to

possibly examine the amount of fringe effects that are present in the devices fabricated

using this mask design. The device level layout for the smaller devices can be seen in

Figure 4.5.

Figure 4.3 (b) shows one of the crossbar designs, the horizontal lines reside in the

top metal mask, and the vertical lines represent the bottom metal. All of the larger pads

are on the top metal mask so all the devices can be accessed using probes. The horizontal

row of pads along the bottom are connected to the lower electrode as the etch mask is

used to provide access. This will allow a voltage to be applied between one top wire and

29
one bottom wire to access any of the devices in the array. Each of the 36 devices at each

of the cross-points is defined by a circle with an area of 100 µm2. The layout in Figure

4.3 (d) is almost identical to the one in Figure 4.3 (c), except that there is no circle added

and the devices are defined only by the crossing of the vertical and horizontal 5 µm wires

for a device area of 25 µm2.

(a) (b)

(c) (d)
Figure 4.3. Layout of 4 different areas in the mask set including (a) the larger individual devices, (b) the
crossbar with 100 µm2 overlap, (c) the smaller individual devices, and (d) the crossbar with 25 µm2
overlap.

30
. . .

Figure 4.4. Span of devices that are included in the larger device set (25 µm2 to 400 µm2).

. . .

Figure 4.5. Span of devices in the smaller devices set (0 µm2 to 35 µm2).

4.2 Second Generation Device Fabrication

After the mask set was designed, it was used to fabricate a wafer for experimental

testing. The device layout can be seen in Figure 4.6. A high resistivity SiO2 film was

deposited on a Si wafer so that the bottom electrode on each device would not short

together. The Pt bottom metal was deposited using electron beam evaporation. Due to the

high melting temperature of platinum, the process had to be completed in steps to prevent

the heat radiated from the Pt target from destroying the wafer. Images of the wafer after

Pt deposition can be seen in Figure 4.7. The TiO2 and TiO2-x layer were then deposited by

sputter deposition using a Ti target with 6% O2 in a 4mTorr Ar environment. The amount

of DC power was varied to determine the composition of each oxide layer (230W for

31
TiO2-x and 50W for TiO2). The stoichiometric oxide layer was set to be much thinner than

the oxygen deficient layer to create a device similar to the one described in [26]. An

image of the wafer with the deposited oxide layers can be seen in Figure 4.8. The Pt top

metal was applied to the wafer again using electron beam evaporation. The photoresist

patterned for the top metal can be seen in Figure 4.9, and the completed wafer can be

seen in Figure 4.10.

Au 500Å

Pt 500Å

Ti 50Å
TiO2 40Å

TiO2-X 1100Å

Pt 500Å
Ti 50Å

SiO2 1mm

Figure 4.6. Device layout for second generation memristor device (contact pads not shown).

(a) (b)
Figure 4.7. Images of the SiO2 wafer after Pt deposition displaying (a) the contact pad and wire leading to
an isolated device, and (b) the wire passing through the circular definition of a crossbar device.

32
Figure 4.8. Image displaying memristor wafer after deposition of TiO2 and TiO2-x layers.

(a) (b)
Figure 4.9. Images displaying the lithography pattern for the top metal for (a) an isolated device, and (b) a
device in a crossbar.

(a) (b)
Figure 4.10. Images of fabricated wafer displaying (a) portion of a memristor crossbar, and (b) an array of
single devices in a wire bonded packaging.

33
4.3 Second Generation Device Characterization

The six plots in Figure 4.11 show the test results for an isolated memristor device

with a cross-sectional area of 200µm2. Tests 1 through 6 were performed in order to

observe how the switching effects of the device change when different voltages were

applied.

In all tests, the voltage sweep went from 0V to the max voltage, from the max

voltage to the minimum voltage, and then from the minimum back to zero. Test 1 was the

initial test on the device, and it appears that the memristor was already in the ‗on‘ state as

the first non-zero data point hit the current limit (about 550µA). The current was limited

until the voltage increase towards 0V. At about -1V, the memristor can be seen switching

to the ‗off‘ state as the conductivity dropped to a non-zero voltage.

Tests 2 and 3 were performed in a similar manner with a lower maximum voltage.

In these tests, the device was not able to switch completely to the on state but a small

hysteresis can be seen. In test 4, the maximum voltage was increased to 4V and the size

of the hysteresis was slightly larger. In these three tests, a negative voltage caused the

device to be conductive but very little hysteresis was displayed.

In test 5, the maximum voltage was pushed to 5V, and this caused a large

switching effect in both the positive and negative regimes. The device went into a very

conductive state after a positive voltage threshold was reached (about 3V), and returned

to the high resistance state as the voltage approached 0 from the negative region (at about

-1V). This result is similar to a device characterization that was published by the

University of Michigan for an a-Si based device [8]. The main difference is that the

conductive state resistance can be calculated when looking at the University of Michigan

34
device, as there are data points present at low voltages (this test also shows a much higher

current limit). Also, the ‗off‘ switching in the device in [8] appears to occur when the

negative voltage is increasing in magnitude. The ‗off‘ switching effect in the device

described in Figure 4.11 appears to occur when the negative voltage is decreasing in

magnitude. Since the switching of a memristor is said to be related to the summation of

the current through the device, a switching that occurs on the voltage increase could still

be valid. Enough current may not have accumulated in the voltage decrease due to the

current limit of the analyzing system.

Test 6 shows a final I-V test when reducing the voltage to see if the result in test 5

had a forming effect on the device. The result is similar to tests 2 – 4 except the

hysteresis is much smaller, and almost negligible relative to the conductivity.

A 100 µm2 device in a crossbar was also tested and the result can be seen in

Figure 4.12. This device has a much lower conductivity, and this could be caused by the

alternate current paths in the crossbar structure. The first test was performed with a max

voltage of 3V and the second had a max voltage of 5V. The size of the hysteresis

increased as the voltage was increased.

35
Test 1 Test 2

500 500

Current (A)

Current (A)
0 0

-500 -500

-5 -4 -3 -2 -1 0 1 2 3 4 5 -5 -4 -3 -2 -1 0 1 2 3 4 5
Voltage (V) Voltage (V)
Test 3 Test 4

500 500
Current (A)

Current (A)
0 0

-500 -500

-5 -4 -3 -2 -1 0 1 2 3 4 5 -5 -4 -3 -2 -1 0 1 2 3 4 5
Voltage (V) Voltage (V)
Test 5 Test 6

500 500
Current (A)

Current (A)

0 0

-500 -500

-5 -4 -3 -2 -1 0 1 2 3 4 5 -5 -4 -3 -2 -1 0 1 2 3 4 5
Voltage (V) Voltage (V)

Figure 4.11. Test results for the 200µm2 memristor device.


Test 1 Test 1
0.1 2

0
0

-2
-0.1
Current (A)

Current (A)

-4
-0.2
-6

-0.3
-8

-0.4
-10

-0.5 -12
-5 -4 -3 -2 -1 0 1 2 3 4 5 -5 -4 -3 -2 -1 0 1 2 3 4 5
Voltage (V) Voltage (V)

Figure 4.12. Test results for the 100µm2 crossbar memristor device.

36
CHAPTER V

COMPARISON OF DEVICE MODELING METHODS

To develop simulations of more complex memristor based systems, the device

models must be studied. If the memristor is to be used to simulate a synapse in

neuromorphic architecture, then it must be capable of holding many different independent

resistance states. Three existing models for the memristor device were studied to

determine the number of resistance levels that can be discerned dependent on device

thickness and the resistance of the device materials.

5.1 Modeling Equations

The equations used to develop the memristor model were introduced in [1], with a

more comprehensive analysis completed in [14]. HP labs described the I-V relationship

for the TiO2 memristor using equation (5.1). This equation states that the memristor can

be modeled using two resistors in series where the resistance of each is dependent

on ( ), the dynamic thickness of the oxygen deficient (low resistivity) layer. For

analysis purposes, ( ) can be seen as the position of a mobile barrier between the doped

and undoped layers. is the resistance of the device when ( ) is equal to 1 and

is the resistance of the device when ( ) is equal to 0; is the total thickness of

both of the oxide layers.

37
( ) ( )
( ) ( ( )) ( ) (5.1)

According to [1], the drift velocity, of the charges in the oxygen deficient layer can be

described using equation (5.2).

( ) (5.2)

The term is equal to ±1 depending on the orientation of the memristor. If

then the width of the oxygen deficient layer will expand under positive bias, otherwise it

will contract. The term represents the mobility of the charges in the low resistivity

layer. Equation (5.2) describes a linear drift velocity for the memristor where the charges

migrate at the same rate regardless of the position of the barrier between the layers. A

more accurate model is described in [14] where a non-linear drift velocity is introduced,

velocity now changes based on position of the barrier between the high and low

resistivity layers. The drift velocity should be at a maximum when the barrier between

the layers is in the middle of the device, and zero when the position of the barrier is at

either end [4]. Joglekar proposed equation (5.3), a window function that is used to

achieve the desired velocity based on barrier position. The term is a positive integer

that controls the intensity at which the function drives velocity to zero at either end of the

device.

( ) ( )
( ) ( ( ) ) (5.3)

Similarly, Biolek et al. proposed another window function in [15] to produce the

same effect with a slightly different shape using equations (5.4) and (5.5).

38
( ) ( )
( ) (( ) ( )) (5.4)

( ) { (5.5)

The value is the current through the memristor. Equation (5.6) shows the drift velocity

definition with the addition of a window function [4-6].

( )
( ) ( ) (5.6)

Recall that equation (1) is a function of ( ) that relates voltage to current. So the

complex relationship shown in equation (5.6) must be solved in terms of ( ) to provide

a current-voltage relationship. A numerical time step solution was developed in [45] to

generate a solution for the current in the memristor based on input voltage. This can be

seen in equations (5.7) to (5.10).

( ) ( )
( ( )) ( ) (5.7)

( )
( ) (5.8)
( ( ))

( )
( ) ( ) ( ) (5.9)

( ) ( )[ ] ( ) (5.10)

In this system of equations, ( ( )) can be calculated assuming ( )

is a known value, which represents the initial thickness of the doped layer. It will also be

assumed that ( ) since no charges in the device are moving when there is no

voltage applied. The next section discusses the implementation of the memristor models

in MATLAB.

39
5.2 Equation Based Device Simulation

Existing studies of these models have utilized sinusoidal inputs to examine if the

output has the characteristic hysteresis seen in memristors. Such inputs cannot be used to

examine the number of resistance levels. Therefore, this study focuses on the application

of successive positive voltage sweeps to determine the number of different resistive states

that can be achieved in the devices. The applied voltage is in the form of a triangular

wave that oscillates between zero and a peak voltage in the range of .25V to 2V. The

initial thickness of the oxygen deficient layer is set to be for all of the simulations.

This is done to provide more room for ( ) to increase while remaining conscious of the

fact that a value any lower may produce results for an unrealistic fabrication. Lastly, is

set equal to 1 for all of the following simulations.

The simulation results discussed first show characterization information based on

the triangular voltage input signal for each of the three models. The three models are

considered to be the linear drift model, Joglekar‘s non-linear drift model, and Biolek‘s

non-linear drift model. Then, as a way to provide theoretical information about the

number of resistance states obtainable, simulations were carried out that display a

relationship between device thicknesses, the ratio , and the number of

achievable resistance states. In the tests conducted, MATLAB was used to test the linear

drift model as well as each of the models using window function to produce non-linear

drift. The next section discusses the results obtained when comparing the models.

5.2.1 Device Simulation Based on Linear Model

The results for the characterization plots of the linear model are shown in Figure

5.1. The plot on the top left shows the voltage and current waveforms for the simulation.

40
The current increases with every triangle pulse that is applied to the device. This is the

intuitive result, as the plot on the bottom right shows that the resistance drops with each

pulse applied. The top right plot shows the I-V characteristic for this simulation. There is

a separate hysteresis loop for each pulse since the variable ( ) is at a different position

at the beginning of each period of the voltage input. When considering the response of

the device when a sine wave is applied, the net charge through the device is zero at the

end of each period. This causes the variable ( ) to be periodic with an overlapping

hysteresis for multiple periods, so no state information is obtained. It can be seen in

Figure 5.1 that the size of the hysteresis increases with each period of the voltage

waveform. This is correlates to the voltage and current waveforms where each current

pulse grows at an exponential rate. The plot on the bottom left shows the thickness of the

doped layer over time as a ratio to the total thickness of the device. It is seen that the

doped layer expanded due to the voltage applied.

Voltage and Current Waveforms I-V Curve


0.3 1.5 1

0.8
Current (mA)
Current (mA)
Voltage (V)

0.2 1
0.6

0.4
0.1 0.5
0.2

0 0 0
0 2 4 6 0 0.05 0.1 0.15 0.2 0.25
time (s) Voltage (V)
w(t) - Width of Doped Region Resistance Change
1
1000
0.8
Resistance ()

800
w(t)/D

0.6 600
0.4 400

0.2 200

0 0
0 2 4 6 0 2 4 6
time (s) time (s)

Figure 5.1. Simulation results for testing liner drift velocity model. The top left plot shows the current and
voltage waveforms with respect to time, and the top right plot shows the I-V characteristic. The bottom left
plot shows position of the barrier between the titanium oxide layers and the bottom right plot shows
resistance as a function of time. Simulation parameters are, RON=100Ω, ROFF=1000Ω, D=46nm, µD=10-
14 2 -1 -1
m V s , w0/D=.2, with a triangle voltage input where f=2Hz and A=.25V.

41
5.2.2 Device Simulation Based on Non-Linear Model 1

Figure 5.2 provides the simulation results when the non-linear drift velocity is

defined by Joglekar‘s window function with . A low value for was chosen

because this limits the last hysteresis loop from increasing dramatically. The current and

voltage waveforms can be seen in the upper left; they look similar to the linear device,

except that the exponential increase in current reduces between the last two pulses. The

hysteresis in the I-V curve on the right shows that the widest loop is now in the middle

where the window function is a maximum. The plots for ( ) and the resistance also

look similar to Figure 5.1, but they have a smoother decay as the window function is

affecting the drift velocity near the edges of the device.

Voltage and Current Waveforms I-V Curve


0.25 2.5 2.5

0.2 2 2
Current (mA)
Current (mA)
Voltage (V)

0.15 1.5 1.5

0.1 1 1

0.05 0.5 0.5

0 0 0
0 2 4 6 0 0.05 0.1 0.15 0.2 0.25
time (s) Voltage (V)
w(t) - Width of Doped Region Resistance Change
1
1000
0.8
Resistance ()

800
w(t)/D

0.6 600
0.4 400

0.2 200

0 0
0 2 4 6 0 2 4 6
time (s) time (s)
Figure 5.2. Simulation results for testing the Joglekar non-linear drift velocity model with p=1. The top left
plot shows the current and voltage waveforms w. r. t. time, and the top right plot shows the I-V
characteristic. The bottom left plot shows position of the barrier between the titanium oxide layers and the
bottom right plot shows resistance as a function of time. Simulation parameters are, RON=100Ω,
ROFF=1000Ω, D=37nm, µD=10-14m2V-1s-1, w0/D=.2, with a triangle voltage input where f=2Hz and A=.25V.

42
5.2.3 Device Simulation Based on Non-Linear Model 2

Figure 5.3 shows the test results when the drift velocity has a non-linearity

induced by Biolek‘s window function. The results look very similar to the data in Figure

5.2, except that the window function has a slightly stronger effect. The value for was

set to 32nm for this test as opposed to 37nm when using Joglekar‘s window function. The

total thickness of the device was reduced slightly to provide an increase in the memory

effect to compensate for the additional reduction provided by the stronger window

function. It can also be seen that this window function only limits the motion at the

boundary to the change in the state variable is tending towards. This can be seen when

comparing the first 2 seconds of the state variable plots in Figures 5.2 and 5.3. It is harder

to see in the I-V plot because not much detail can be seen when resistance is near the

maximum (ROFF).

Voltage and Current Waveforms I-V Curve


0.25 2.5 2.5

0.2 2 2
Current (mA)
Current (mA)
Voltage (V)

0.15 1.5 1.5

0.1 1 1

0.05 0.5 0.5

0 0 0
0 2 4 6 0 0.05 0.1 0.15 0.2 0.25
time (s) Voltage (V)
w(t) - Width of Doped Region Resistance Change
1
1000
0.8
Resistance ()

800
w(t)/D

0.6 600
0.4 400

0.2 200

0 0
0 2 4 6 0 2 4 6
time (s) time (s)
Figure 5.3. Simulation results for testing the Biolek non-linear drift velocity model with p=1. The top left
plot shows the current and voltage waveforms w. r. t. time, and the top right plot shows the I-V
characteristic. The bottom left plot shows position of the barrier between the titanium oxide layers and the
bottom right plot shows resistance as a function of time. Simulation parameters are, RON=100Ω,
ROFF=1000Ω, D=32nm, µD=10-14m2V-1s-1,w0/D=.2, with a triangle voltage input where f=2Hz and A=.25V.

43
5.3 Resistance State Analysis Based on Device Models

When implementing a neuromorphic circuit, it is necessary to know how many

different resistance levels are available in a given memristor device. The models that

have been discussed were used to generate data that predicts the number of resistance

states available in a device based on different device parameters. Since this is an analog

device, the definition of a resistance level will not be an on/off relationship. Additionally,

the exponential nature of the resistance drop must be considered when defining a level.

The amplitude and frequency of the input voltage will also have an effect the number of

levels; these however were held constant in the experiments to isolate how device

parameters determine the resistance levels available in the device.

Since the resistance drop is not constant for each pulse applied, a resistance level

has been defined as any state change in ( ) that causes a sufficient increase in current

upon the application of a single voltage pulse. The sufficient increase in current was

given as a percentage increase from the previous current pulse. When looking at the

current and voltage waveforms in Figure 5.3, it can be seen that the first four pulses

hardly changed the device state, and the last four pulses had a much larger impact. The

tolerance percentage decided at what value of ( ) the resistance level changed upon a

single voltage pulse.

5.3.1 Resistance States in the Linear Device Model

Figure 5.4 shows the data collected for testing the linear drift model. The plot on

the left shows the number of resistance levels available as a function of device thickness

when the ratio . The plot on the right shows the same data when

44
. Since no windowing function was employed with this model, the starting

thickness was chosen by finding the minimum whole nanometer thickness where ( )

. In Figure 5.4 it can be seen that tolerances of 5%, 10%, and 15% were considered

as three separate tolerance levels for the increase in peak value for current. This data

relates to the uniformity at which the devices are fabricated. If all of the memristor

devices in a system are very uniform then a smaller current peak differential is required

because the same magnitude voltage pulse is likely to change the resistance of all the

devices by the same amount. If non-uniform devices are fabricated, a larger gap is needed

between resistance levels because internal resistance may change dramatically from one

device to the next.

Linear Memristor Model, Roff/Ron=10 Linear Memristor Model, Roff/Ron=50


9 9
Min Pulse Differental 5% Min Pulse Differental 5%
Min Pulse Differental 10% Min Pulse Differental 10%
8 Min Pulse Differental 15% 8 Min Pulse Differental 15%
Number of Resistance Levels Availiable

Number of Resistance Levels Availiable

7 7

6 6

5 5

4 4

3 3

2 2

1 1

0 0
90 95 100 105 110 115 40 45 50 55
D (nm) D (nm)

Figure 5.4. Plot on the left shows number of resistance levels achieved using the linear drift model for
D=90-115nm where RON/ROFF=10 for current pulse differentials given as 5, 10, and 15%. Plot on the right
shows same data for RON/ROFF=50 and D=42-55nm. Other simulation parameters are, RON=100Ω, µD=10-
14 2 -1 -1
m V s , w0/D=.2, with a triangle voltage input where f=4Hz and A=2V.

5.3.2 Resistance States in the Non-Linear Device Models

Figure 5.5 shows the resistance level results for the simulation when using

Joglekar‘s non-linear drift model. The main difference here was that the number of levels

increased initially before the expected decay. This is because at these small thicknesses,

the memristive effect was so large that very few pulses were needed to drive ( )
45
Once ( ) reaches this limit, a negative pulse must be applied or the device will remain

in this state as a linear resistor. The data in Figure 5.6 shows the simulation results for a

case where ( ) so quickly that the number of resistance levels was

reduced. The bottom left plot in Figure 5.6 shows that after about 1.75s, there was no

more room for the oxygen deficient layer to expand and the memristor started to act as a

linear resistor. At this time, the peak current no longer increased with the application of

additional pulses and the I-V characteristic became a straight line with slope equal to

Figure 5.5 shows that when the device thickness was less than 65nm, this effect

occurred as the number of resistance levels decreased as device thickness decreased.

Lastly, Figure 5.7 shows the resistance level data collected for the Biolek window

function. The results are similar to the Joglekar model but the span of device thicknesses

was slightly reduced. Since the window function for the Biolek model had a greater effect

on drift velocity, the maximum number of resistance levels occurred at a lower device

thickness.

Joglekar Memristor Model (p=1), Roff/Ron=10 Joglekar Memristor Model (p=1), Roff/Ron=50
10 10 Min Pulse Differental 5%
Min Pulse Differental 5%
Min Pulse Differental 10% Min Pulse Differental 10%
9 9
Min Pulse Differental 15% Min Pulse Differental 15%
Number of Resistance Levels Availiable

Number of Resistance Levels Availiable

8 8

7 7

6 6

5 5

4 4

3 3

2 2

1 1

0 0
60 80 100 120 30 35 40 45 50
D (nm) D (nm)

Figure 5.5. Plot on the left shows number of resistance levels achieved using the Joglekar non-linear drift
model for D=60-120nm where RON/ROFF=10 for current pulse differentials given as 5, 10, and 15%. Plot on
the right shows same data for RON/ROFF=50 and D=30-50nm. Other simulation parameters are, RON=100Ω,
µD=10-14m2V-1s-1, w0/D=.2, with a triangle voltage input where f=4Hz and A=2V.

46
Voltage and Current Waveforms I-V Curve
2 20 20

1.5 15 15

Current (mA)
Current (mA)
Voltage (V)
1 10 10

0.5 5 5

0 0 0
0 1 2 3 0 0.5 1 1.5 2
time (s) Voltage (V)
w(t) - Width of Doped Region Resistance Change
1 1000
0.8

Resistance ()
800
w(t)/D

0.6 600
0.4 400

0.2 200

0 0
0 1 2 3 0 1 2 3
time (s) time (s)

Figure 5.6. Simulation results for testing the Joglekar non-linear drift velocity model with p=1. The top left
plot shows the current and voltage waveforms w. r. t. time, and the top right plot shows the I-V
characteristic. The bottom left plot shows position of the barrier between the titanium oxide layers and the
bottom right plot shows resistance as a function of time. Due to the strong memristive effect, device
operates as a linear resistor after 1.5s. Simulation parameters are, RON=100Ω, ROFF=1000Ω, D=50nm,
µD=10-14m2V-1s-1,w0/D=.2, with a triangle voltage input where f=4Hz and A=2V.

Biolek Memristor Model (p=1), Roff/Ron=10 Biolek Memristor Model (p=1), Roff/Ron=50
12 Min Pulse Differental 5% 12 Min Pulse Differental 5%
Min Pulse Differental 10% Min Pulse Differental 10%
Min Pulse Differental 15% Min Pulse Differental 15%
10 10
Number of Resistance Levels Availiable

Number of Resistance Levels Availiable

8 8

6 6

4 4

2 2

0 0
50 60 70 80 90 100 110 30 35 40 45 50
D (nm) D (nm)

Figure 5.7. Plot on the left shows number of resistance levels achieved using the Biolek non-linear drift
model for D=50-110nm where RON/ROFF=10 for current pulse differentials given as 5, 10, and 15%. Plot on
the right shows same data for RON/ROFF=50 and D=30-50nm. Other simulation parameters are, RON=100Ω,
µD=10-14m2V-1s-1, w0/D=.2, with a triangle voltage input where f=4Hz and A=2V.

5.3.3 Resistance State Comparison

When comparing the number of resistance levels available for the three models, it

can be seen the two non-linear models provided similar results due to boundaries added

47
to reduce drift velocity at the borders. Figure 5.8 shows the results for all three models

with a pulse differential of 5%. The Biolek and Joglekar models show a similar pattern in

the number of resistance levels, but the Biolek model on average provided about one

more level relative to device thickness. The capability of the linear model suffers because

it immediately turns into a linear resistor when ( ) and it will not operate in the

range where the non-linear models provide the maximum number of resistance levels. In

the range the linear model operates correctly, it fit the pattern of the non-linear models

and provided a slight increase in the number of resistance levels.

Looking at the characterization plots, it can be seen that hysteresis size increases

exponentially with each pulse applied to the device. This is because the resistance drop is

linear with respect to the thickness of the doped layer. When the current is calculated

using equation (5.1), it is inversely proportional to the resistance value producing an

asymptotic increase as ( ) approaches This is displayed in Figure 5.9. As the

thickness of the doped layer approached 0 nm, the conductance of the memristor

increased asymptotically. Since this occurs very close to the boundary of the memristor,

the window functions described in [14, 15] reduce this effect.

When comparing these results to the experimental results in [6, 7, and 10], it can

be seen that the exponential increase in hysteresis and peak current is not present in

published device characterization data. The hysteresis was initially large, and then

diminished with each pulse. In addition, it was shown in [7] that the current pulses

increase logarithmically with each voltage pulse as opposed to exponentially.

This method for determining the number of resistance levels can still be used

since there is still non-linear increase in current. Thus applying this technique to the

48
device described in [7] will show that higher differential current pulses occur when

( ) is closer to 0 as opposed to when ( )

Comparison of Models, Roff/Ron=10, Pulse Diff. = 5% Comparison of Models, Roff/Ron=50, Pulse Diff. = 5%
11 11
Linear Model Linear Model
Joglekar Model Joglekar Model
10 10
Biolek Model Biolek Model

Number of Resistance Levels Availiable

Number of Resistance Levels Availiable


9
9

8
8
7
7
6
6
5

5
4

3 4

2 3
40 60 80 100 120 30 35 40 45 50 55
D (nm) D (nm)

Figure 5.8. Comparison of results for resistance level data presented in previous section using a minimum
pulse differential of 5%.

Relationship Between Thickess of Ti0 2-X Layer and Conductance


0.01
Roff/Ron = 10
Roff/Ron = 100
0.008 Roff/Ron = 200

0.006
1/M(w(t))

0.004

0.002

0
0 0.2 0.4 0.6 0.8 1
w(t)/D
Figure 5.9. Device conductance as a function of doped layer thickness when using simulations based on
equation (1).

When comparing the results from all three models in Figure 5.8, it can be seen

that the linear model could not operate in the range where the non-linear models provided

the maximum number of resistance levels. Also, the Biolek model provided a slightly

higher number of levels than the Joglekar model. When looking at the linear model in its

defined region, it provided more resistance levels on average, but it did not provide the

49
maximum number of levels. In all three cases, increasing the ratio reduced the

thickness where the maximum number of resistance levels was observed.

Looking at the resistance level data, an argument could be made stating that much

of the functionality of the memristor is wasted in the areas where the current peak

differential is low. Furthermore, several pulses could be used to change the resistance

state in less sensitive areas to increase the number of resistance levels. This would require

a more complex circuit that would apply a different number or size of voltage pulses

dependent on the device state. As an alternative for increasing the amount of resistance

levels available, circuit designs could be examined that use multiple memristors to model

a single synapse.

5.4 Device Modeling Based on Fabrication Data

The device models described in the previous section do not contain the threshold

property seen in physical memristor devices [6, 7, 10, and 25]. The threshold is defined

as the minimum voltage that is required to change the state of the memristor device. The

physical properties that determine the threshold of the devices are complex and appear to

be different for many published devices. A model that is capable of providing a

programming threshold was made using the data collected from a physical device

characterized in [10].

A simulation was constructed in MATLAB that tests the developed model with a

string of voltage pulses. The data in [10] shows the device‘s reaction to five positive

voltage sweeps, each represented by a separate hysteresis in the I-V curve. The threshold

voltage for each state can also be seen in the characterization data where overlapping

50
lines begin to diverge in the I-V curve. Since five separate loops were displayed in [10],

each one is considered to be the operating region of a resistance state. The simulation is

capable of determining the memristor device state and the resistance of the device

dependent on the voltage applied and the state the device is in.

Figure 5.10 (a) shows the shows the voltage input as a series of pulses generated

with random amplitude and polarity. The voltage amplitude was generated in a range

from 0 to 0.3V, enough for the voltage to be slightly larger than the highest threshold

point according to the data in [10]. Since the data used to generate the simulation only

was available for positive voltage sweeps, it was assumed that a negative pulse affected

the device state in the same way, just in the opposite direction. Figure 5.10 (b) shows the

state of the memristor device corresponding to the voltage signal applied to the device. It

can be seen that not all pulses changed the state of the device because not all of them

reached the threshold voltage. Also, when looking at specific pulses, it can be seen that

the different states require different pulse amplitudes to increase or decrease the state of

the device. For example, when looking at the first pulse on the left and the third pulse

from the right, they were about the same amplitude, but only the pulse on the left changed

the state of the device. This is because the threshold required to move from state 5 to 4 is

greater than the threshold required to move from state 3 to 2. Figure 5.10 (c) shows the

resistance of the device over time. In general, the resistance increased as the state

decreased, where slight changes in the slope of the I-V curve correspond to the noise in

the resistance plot. It should also be noticed that as the state number increased, the

resistance change became smaller, with a negligible resistance differential between states

4 and 5.

51
Input Voltage to Memristor Model
0.3

0.2

0.1
Amplitude (V)

-0.1

-0.2

0 5 10 15 20 25 30 35 40
time (s)
(a)
State of Memristor Based on Voltage Pulses

4.5
State (1 to 5)

3.5

2.5

2
0 5 10 15 20 25 30 35 40
time (s)
(b)
Dynamic Resistance of Memristive Device
700

600
Resistance ( )

500

400

300

200
0 5 10 15 20 25 30 35 40
time (s)
(c)
Figure 5.10. Simulation results for numerically based model that incorporates programming threshold
displaying (a) the voltage input, (b) the memristor state, and (c) the dynamic resistance value.

52
CHAPTER VI:

CIRCUIT DESIGN BASED ON NON-LINEAR DRIFT MODEL

To further model the behavior of the memristor device, circuits have been

simulated that can use voltage pulses to either change, or sense the state of a device. In

this application the memristor is meant to simulate a synapse, so it is ideal that the

memristor can store a number of different resistances. It is also desirable that these

devices will be compatible with CMOS, both in fabrication and circuit design, so

standard 5V rectangular pulses are used as the input waveforms.

6.1 Circuit for Writing to Memristor

6.1.1 Write Circuit Schematic

The proposed memristor write circuit can be seen in Figure 6.1. The circuit has

two pulse inputs and requires a 5V DC supply to operate. When a pulse is applied from

the source V1, MOSFETs M1 and M2 are turned on and current can flow through the

device from the positive to the negative terminal. This creates a positive voltage drop

across the device which causes the state variable to shift upwards, thus decreasing the

resistance. When a pulse is generated at V2, MOSFETs M3 and M4 are turned on and a

negative voltage drop is formed between the positive and negative terminals of the

memristor. This causes the state variable to decrease, lowering the resistance of the

53
memristor. This technique for programming the memristor is useful for developing spike-

based neural system, as each pulse changes the synaptic weight represented by the

memristor. Furthermore, the artificial neuron only needs to be able to generate positive

voltage pulses for this design.

Figure 6.1 MOSFET based memristor write circuit.

6.1.2 Simulation Results for Memristor Write Circuit

The data in Figure 6.2 shows the simulation results for a train of 5V pulses being

applied to the circuit from V1. It can be seen that the state variable w(t) shifts upward

slightly with each pulse applied, and retains its position when the pulsed waveform is

equal to 0. Figure 6.3 shows a second simulation where the pulses are applied to V2, and

the opposite result is obtained.

54
Voltage Drop Across Memristor Current Through Memristor
100 15

80

Voltage (mV)
10

Current (A)
60

40
5
20

0 0
0 5 10 15 0 5 10 15
time (s) time (s)

0.8
w(t)/D

0.6

0.4
0 5 10 15
time (s)

Figure 6.2. Memristor response to write circuit simulation with voltage applied from V1. Top left plot
shows the voltage across the memristor, when the 5V pulses are applied, the top right plot shows the
current in the memristor and the voltage pulse are applied, and the bottom plot shows the shift if the state
variable w(t) relative to the thickness of the device.

Voltage Drop Across Memristor Current Through Memristor


0 0

-20 -1
Voltage (mV)

Current (A)

-40 -2

-60 -3

-80 -4

-100 -5
0 5 10 15 0 5 10 15
time (s) time (s)

0.5

0.4

0.3
w(t)/D

0.2

0.1

0
0 5 10 15
time (s)

Figure 6.3. Memristor response to write circuit simulation with voltage applied from V2. Top left plot
shows the voltage across the memristor, when the 5V pulses are applied, the top right plot shows the
current in the memristor and the voltage pulse are applied, and the bottom plot shows the shift if the state
variable w(t) relative to the thickness of the device.

55
6.2 Circuit for Reading Resistance of a Memristor Device

6.2.1 Read Circuit Schematic

A circuit was also implemented to read the state of the memristor device. This

was done by using an op-amp as a comparator that has a reference voltage set by a static

resistance. The voltage at the negative terminal is compared to the output voltage of the

memristor, and the comparator output determines the state of the memristor. The read

circuit displayed in Figure 6.4 has two comparators. This means that three different states

can be determined from the memristor. This circuit can be extrapolated to show that the

number of resistance levels is equal to one plus the number of comparators. This is true

as long as the voltage differential between the comparators is greater than the noise in the

circuit or the variation in the memristor devices.

56
Figure 6.4 Comparator based read circuit for memristor.

6.2.2 Simulation Results for Memristor Read Circuit

Figure 6.5 shows the results of a simple simulation where pulses were applied to

the device, and the comparator output was examined to determine the state of the

memristor. When comparing the voltage of the memristor to the voltage at the

comparator output, it can be seen that the comparator with the 2.5V reference is on, but

the comparator with the 4V reference is off.

One of the problems with this circuit is that the resistance of the memristor

changes as the read voltage is applied to the device. Looking at the center plot in Figure

6.5, the voltage is slightly increased with every pulse applied. If the memristor were to be

read many times in a row, the comparator with the 4V reference would eventually turn

on, changing the state of the device. This is problem is caused by the use of a device

57
model that does not have a programming threshold. This should not be a problem when

using physical devices since many of them contain programing thresholds. No

information would be lost is the read pulse magnitude did not exceed the programming

threshold for each device. This could be tested in simulation using a model such as the

one proposed in section 5.4, although this model was not programmed in SPICE.

Read Circuit for Memristor

5
Input Voltage (V)

0
0 1 2 3 4 5 6 7
time (s)
Voltage at Memristor Output (V)

0
0 1 2 3 4 5 6 7
time (s)
Comparator Outputs (V)

5 2.5V Comparator Output


4 4V Comparator Output

0
0 1 2 3 4 5 6 7
time (s)
Figure 6.5. Read circuit simulation. The top plot shows the input voltage, the middle plot shows the voltage
generated between the memristor and the 75k resistor, and the bottom plot shows the output of each of the
comparators.

58
CHAPTER VII

CONCLUSION

The initial steps toward developing memristor based systems have been

completed and the results have been demonstrated in the previous chapters. Memristor

devices have been fabricated and characterized, as these are necessary steps toward

developing a robust and reliable fabrication technique. The existing published memristor

device models have been compared and an alternate modeling method has been proposed

that incorporates a programming threshold. Finally, circuits have been developed that

provide the capability to read from and write to memristors.

The data collected from the characterization tests show that the fabricated devices

display a clear hysteresis loop with a zero crossing, characteristic of memristors. For the

most part, the devices were only conductive when the voltage applied was positive, the

exception being UDMEM2 when the Keithley apparatus was used for testing. When

multiple voltage loops were applied in succession, the hysteresis loops started to degrade

after an initial increase in size for both of the test setups. The UDMEM8 device was

tested most successfully, and possessed a greater conductivity when positively biased.

Furthermore, the resistance of the UDMEM2 device was successfully lowered with the

application of voltage pulses, showing promise for the device when modeling neural

functions.

59
The second generation of memristor devices was based on a more complex mask

set that provided a variable device area with the addition of crossbar arrays. The results

from these devices showed a larger hysteresis in some cases. When the conductivity of

the device was asymmetric, the negative region was observed to be more conductive.

This was most likely due to the fact that the TiO2-x layer was below the TiO2 layer,

leading to a device that was reversed when compared to the first generation devices.

The comparison of the existing device models provided information as to how

memristors may react based on different device thicknesses and material resistances.

When comparing the device model simulations to published memristor devices, there was

a discrepancy in the I-V characteristics in the form of an incorrectly shaped hysteresis.

The window functions that were implemented reduce this effect in the extreme case

where ( ) is close to the boundaries, but an exponential increase can still be observed.

The resistance of the fabricated devices published in [7, 25] appears to change

with respect to the magnitude of the voltage input. This may be caused by the metal-

insulator-metal (MIM) structure of memristors, which may introduce a dynamic

resistance component dependent on magnitude of the voltage input [33]. However, the

device models that were compared do not contain a conductivity component dependent

on the magnitude of the input voltage.

Furthermore, the published device characterizations [6, 7, and 10] show a

programming threshold property that was not present in the existing models. As a data

driven solution to this problem, a model was developed directly based on the

characterization published in [10].

60
The write circuit is capable of increasing or decreasing the resistance of a

memristor using a sequence of positive voltage pulses applied to different inputs. The

simulations confirmed that the circuit was operating properly although the process was

slow compared to alternative memory structures. The read circuit was able to determine

the quantized state of a memristor based on a set of comparator blocks. The simulation

showed that the circuit can function although the read pulse is destructive and could

change the state of the memristor over a large series of reads. This would not be a

problem if the memristor device model contained a threshold property and the read pulse

did not exceed the programming threshold.

Future work includes developing memristor devices that operate more predictably

and determining a more reliable fabrication method. Also a more advanced memristor

model will be developed that is capable of modeling published characterization data

much more accurately. This will provide the ability to develop much more accurate

circuit simulations and power analysis based on different types of memristor devices.

Future circuit designs will incorporate memristor arrays to show how these read and write

methods can be applied to several memristors either sequentially or in parallel. These

circuits will then be used to develop larger scale circuit simulations for determining the

performance improvement when utilizing memristors for memory systems or

neuromorphic architectures.

61
BIBLIOGRAPHY

[1] D. B. Strukov, G. S. Snider, D. R. Stewart, and R. S. Williams, ―The missing


Memristor found,‖ Nature, 453, 80–83 (2008).

[2] C. Yakopcic, T. M. Taha, G. Subramanyam, E. Shin, P. T. Murray, and S. Rogers,


―Memristor-Based Pattern Recognition for Image Processing: an Adaptive Coded
Aperture Imaging and Sensing Opportunity,‖ SPIE Adaptive Coded Aperture
Imaging and Non-Imaging Sensors. (2010).

[3] C. Yakopcic, E. Shin, T. M. Taha, G. Subramanyam, P. T. Murray, and S. Rogers,


―Fabrication and Testing of Memristive Devices,‖ International Joint Conference
on Neural Networks (IJCNN). (2010).

[4] C. Yakopcic, T. M. Taha, E. Shin, G. Subramanyam, P. T. Murray, and S. Rogers,


―Memristor Fabrication and Characterization; An Adaptive Coded Aperture
Imaging & Sensing Opportunity,‖ SPIE Adaptive Coded Aperture Imaging and
Non-Imaging Sensors. (2010).

[5] L. O. Chua, Leon O, "Memristor—The Missing Circuit Element," IEEE


Transactions on Circuit Theory, 18(5), 507–519 (1971).

[6] G. S. Snider, ―Cortical Computing with Memristive Nanodevices,‖ SciDAC Review,


(2008).

[7] S. H. Jo, T. Chang, I. Ebong, B. B. Bhadviya, P. Mazumder, and W. Lu, ―Nanoscale


Memristor Device as Synapse in Neuromorphic Systems,‖ Nano Letters, 10 (2010).

[8] S. H. Jo, and W. Lu, ―CMOS Compatible Nanoscale Nonvolatile Resistance


Switching Memory,‖ Nano Letters, 8(2), 392-397 (2008).

[9] R. E. Pino, J. W. Bohl, N. McDonald, B. Wysocki, P. Rozwood, K. A. Campbell,


A. Oblea, and A. Timilsina, ―Compact Method for Modeling and Simulation of
Memristor Devices: Ion conductor chalcogenide-based memristor devices,‖
International Symposium on Nanoscale Architectures, (2010).

[10] A. S. Oblea, A. Timilsina, D. Moore, K. A. Campbell, ―Silver Chalcogenide Based


Memristor Devices,‖ International Joint Conference on Neural Networks (IJCNN),
(2010).

62
[11] N. Gergel-Hackett, B. Hamadani, B. Dunlap, J. Suehle, C. Richter, C. Hacker, D.
Gundlach, ―A Flexible Solution-Processed Memristor,‖ IEEE Electron Device
Letters, 30(7), (2009).

[12] T. Prodromakis, K. Michelakis, C. Toumazou, ―Switching Mechanisms in


Microscale Memristors,‖ Electronics Letters 46(1), (2010).

[13] Prodromakis, K. Michelakis, C. Toumazou, ―Fabrication and Electrical


Characteristics of Memristors with TiO2/TiO2+X Active Layers,‖ IEEE International
Symposium on Circuits and Systems, (2010).

[14] Y. N. Joglekar and S. J. Wolf, "The elusive memristor: properties of basic electrical
circuits," European Journal of Physics, 30, 661–675 (2009).

[15] Z. Biolek, D. Biolek, V. Biolková, "Spice Model of Memristor with Nonlinear


Dopant Drift," Radioengineering, 18(2), 210-214 (2009).

[16] Y. Ho, G. M. Huang, P. Li, ―Dynamical Properties and Design Analysis for
Nonvolatile Memristor Memories,‖ IEEE Transactions on Circuits and Systems I:
Regular Papers, (2010).

[17] D. Niu, Y. Chen, Y. Xie, ―Low-power Dual-element Memristor Based Memory


Design,‖ ACM/IEEE International Symposium on Low-Power Electronics and
Design, (2010).

[18] Y. Ho, G. M. Huang, P. Li, ―Nonvolatile Memristor Memory: Device


Characteristics and Design Implications,‖ IEEE/ACM International Conference on
Computer-Aided Design – Digest of Technical Papers, (2009).

[19] H. Kim, M. P. Sah, C. Yang, L. O. Chua, ―Memristor-based Multilevel Memory,‖


12th International Workshop on Cellular Nanoscale Networks and Their
Applications (CNNA), (2010).

[20] E. Lehtonen, M. Laiho, ―CNN Using Memristors for Neighborhood Connections,‖


12th International Workshop on Cellular Nanoscale Networks and Their
Applications (CNNA), (2010).

[21] M. Laiho, E. Lehtonen, ―Cellular Nanoscale Network Cell with Memristors for
Local Implication Logic and Synapses,‖ IEEE International Symposium on Circuits
and Systems, (2010).

[22] L. O. Chua, S. M. Kang, ―Memristive Devices and Systems,‖ Proceedings of the


IEEE, 64(2), 1976.

[23] R. S. Williams, ―How We Found The Missing Memristor,‖ IEEE Spectrum, vol. 45,
no. 12, pp. 28-35, 2008.

63
[24] G. S. Snyder, ―Spike-timing-dependent learning in memristive nanodevices,‖ IEEE
International Symposium on Nanoscale Architectures, 85-92 (2008).

[25] J. J. Yang, M. D. Pickett, X. Li, D. A. A. Ohlberg, D. R. Stewart and R. S.


Williams, ―Memristive switching mechanism for metal/oxide/metal nanodevices,‖
Nature Nanotechnology, 3, 429–433 (2008).

[26] M. D. Pickett, D. B. Strukov, J. L. Borghetti, J. J. Yang, G. S. Snider, D. R. Stewart,


and R. S. Williams, ―Switching dynamics in titanium dioxide memristive devices,‖
Journal of Applied Physics 106(074508), (2009).

[27] J. Borghetti, D. B. Strukov, M. D. Pickett, J. J. Yang, D. R. Stewart, R. S. Williams,


―Electrical transport and thermometry of electroformed titanium dioxide memristive
switches.‖ Journal of Applied Physics 106(124504), (2009).

[28] F. Miao, J. J. Yang, J. P. Strachan, D. Stewart, R. S. Williams, C. N. Lau, ―Force


modulation of tunnel gaps in metal oxide memristive nanoswitches,‖ Applied
Physics Letters, 95(113503), (2009).

[29] F. Miao, D. Ohlberg, D. R. Stewart, R. S. Williams, C. N. Lau, ―Quantum


Conductance Oscillations in Metal/Molecule/Metal Switches at Room
Temperature,‖ Physical Review Letters, 101(016802), (2008).

[30] J. P. Strachan, J. J. Yang, R. Munstermann, A. Scholl, G. Medeiros-Ribeiro, D. R.


Stewart, R. S. Williams, ―Structural and chemical characterization of TiO2
memristive devices by spatially-resolved NEXAFS,‖ Nanotechnology, 20(485701),
(2009).

[31] J. J. Yang, F. Miao, M. D. Pickett, D. A. A. Ohlberg, D. R. Stewart, C. N. Lau, R.


S. Williams, ―The mechanism of electroforming of metal oxide metal memristive
switches,‖ Nanotechnology, 20(215201), (2009).

[32] J. P. Strachan, D. B. Strukov, J. Borghetti, J. J. Yang, G. Medeiros-Ribeiro, and R.


S. Williams, ―The switching location of a bipolar memristor: chemical, thermal and
structural mapping,‖ Nanotechnology, 22(254015), (2011).

[33] J. G. Simmons, ―Generalized Formula for the Electric Tunnel Effect between
Similar Electrodes Separated by a Thin Insulating Film,‖ Journal of Applied
Physics, 34(6), (1963).

[34] D. H. Kwon, K. M. Kin, J. H. Jang, J. M. Jeon, M. H. Lee, G. H. Kim, X. S. Li, G.


S. Park, B. Lee, S. Han, M. Kim, C. S. Hwang, ―Atomic structure of conducting
nanofilaments in TiO2 resistive switching memory,‖ Nature Nanotechnology,
(2010).

[35] X. Wang, Y. Chen, ―Spintronic Memristor Devices and Application,‖ Design,


Automation & Test in Europe Conference and Exhibition, (2010).

64
[36] X. Wang, Y. Chen, H. Xi, H. Li, D. Diminov, ―Spintronic Memristor Through Spin-
Torque-Induced Magnetization Motion,‖ IEEE Electron Device Letters, 30(3),
2009.

[37] Y. Chen, X. Wang, ―Compact Modeling and Corner Analysis of Spintronic


Memristor,‖ IEEE/ACM International Symposium on Nanoscale Architectures,
(2009).

[38] X. P. Wang, Y. Y. Chen, L. Pantisano, L. Goux, M. Jurczak, G. Groeseneken, and


D. J. Wouters, ―Effect of Anodic Interface Layers on the Unipolar Switching of
HfO2-based Resistive RAM,‖ International Symposium on VLSI Technology
Systems and Applications, 140-141 (2010).

[39] A. S. Oblea, R. E. Pino, J. W. Bohl, A. Timilsina, K. A. Campbell, ―Memristor


SPICE Model Simulation & Device Hardware Correlation,‖ (2010).

[40] W. Wang, Q. Yu, C. Xu, Y. Cui, ―Study of Filter Characteristics Based on PWL
Memristor,‖ International Conference on Communications, Circuits, and Systems,
(2009).

[41] S. Shin, K. Kim, S. M. Kang, ―Memristor Applications For Programmable Analog


ICs,‖ IEEE Transactions on Nanotechnology, (2010).

[42] K. Eshraghian, K. R. Cho, O. Kavehei, S. K. Kang, D. Abbott, S. M. Kang,


―Memristor MOS Content Addressable Memory (MCAM): Hybrid Architecture for
Future High Performance Search Engines,‖ IEEE Transactions on Very Large
Integration (VLSI) Systems, (2010).

[43] S. H. Jo, K. H. Kim, and W. Lu, ―High-Density Crossbar Arrays Based on a Si


Memristive System,‖ Nano Letters, 9(2), (2009).

[44] A. Afifi, A. Ayatollahi, F. Raissi, ―Implementation of Biologically Plausible


Spiking Neural Network Models On the Memristor Crossbar based CMOS/Nano
Circuits,‖ European Conference on Circuit Theory and Design, (2009).

[45] N. R. McDonald, R. E. Pino, P. J. Rozwood, and B. T. Wysocki, ―Analysis of


Dynamic Linear and Non-linear Memristor Device Models for Emerging
Neuromorphic Computing Hardware Design,‖ IEEE World Congress of
Computational Intelligence, (2010).

[46] T. Raja, S. Mourad, ―Digital Logic Implementation in Memristor-based Crossbars,‖


Fifth IEEE Symposium on Electronic Design, Test and Application, (2010).

65

You might also like