ThesisPDF Available

Abstract and Figures

Neural activity recording system promotes the development of diagnostic and therapeutic programs and neuroscience research. Direct recordings of neural signals from the brain have helped scientists access to study and unlock the secrets of neural coding gradually. This can be realised by applying implantable neural recording systems to monitor and record neural signals. Then, the neural information can be transmitted to the external device for processing, storage or application. However, the power consumption of the neural recording system is the primary constraint to monitoring large groups of neurons. It leads the development of neural recording systems in two directions: 'high-channel-count but wired' and 'wireless but low-channel-count'. To address the power issue, we proposed a neural front-end that aims to detect neural spikes by thresholding and output as one-bit digital data so that the afterwards processing can only work on spikes rather than processing all the data points. The most significant feature is that we induce memristors as trimming devices to tune the threshold voltage for spike detection. Meanwhile, it contributes to rejecting up to 50mV DC offset from electrodes. The measurement presents that the memristor-based pre-amplifier is capable of achieving above 95% spike detection accuracy with hundreds of nanowatt power consumption per channel. This design indicates a promising approach to conduct spike-detection on-chip with low power consumption and demonstrates the potential of a hybrid memristor/CMOS circuit for power-efficient large-scale neural interfacing application.
Content may be subject to copyright.
University of Southampton Research Repository
Copyright ©and Moral Rights for this thesis and, where applicable, any accompanying data are
retained by the author and/or other copyright owners. A copy can be downloaded for personal non-
commercial research or study, without prior permission or charge. This thesis and the accompanying
data cannot be reproduced or quoted extensively from without first obtaining permission in writing
from the copyright holder/s. The content of the thesis and accompanying research data (where appli-
cable) must not be changed in any way or sold commercially in any format or medium without the
formal permission of the copyright holder/s.
When referring to this thesis and any accompanying data, full bibliographic details must be given, e.g.
Thesis: Jiaqi Wang (2023) ”Ultra-fine Signal Classification Using Memristor-Enabled Hardware”,
University of Southampton, Faculty of Engineering and Physical Sciences, Electronics and Computer
Science, PhD Thesis, 109.
UNIVERSITY OF SOUTHAMPTON
Faculty of Engineering and Physical Sciences
Electronics and Computer Science
Ultra-Fine Signal Classification Using
Memristor-Enabled Hardware
by
Jiaqi Wang
https://orcid.org/0000-0002-2503-1001
A thesis for the degree of
Doctor of Philosophy
February 2023
UNIVERSITY OF SOUTHAMPTON
ABSTRACT
FACULTY OF ENGINEERING AND PHYSICAL SCIENCES
ELECTRONICS AND COMPUTER SCIENCE
Doctor of Philosophy
ULTRA-FINE SIGNAL CLASSIFICATION USING
MEMRISTOR-ENABLE HARDWARE
by Jiaqi Wang
Neural activity recording system promotes the development of diagnostic and therapeutic
programs and neuroscience research. Direct recordings of neural signals from the brain
have helped scientists access to study and unlock the secrets of neural coding gradually.
This can be realised by applying implantable neural recording systems to monitor and
record neural signals. Then, the neural information can be transmitted to the external
device for processing, storage or application. However, the power consumption of the
neural recording system is the primary constraint to monitoring large groups of neurons.
It leads the development of neural recording systems in two directions: ‘high-channel-
count but wired’ and ‘wireless but low-channel-count’. To address the power issue, we
proposed a neural front-end that aims to detect neural spikes by thresholding and out-
put as one-bit digital data so that the afterwards processing can only work on spikes
rather than processing all the data points. The most significant feature is that we in-
duce memristors as trimming devices to tune the threshold voltage for spike detection.
Meanwhile, it contributes to rejecting up to 50mV DC offset from electrodes. The mea-
surement presents that the memristor-based pre-amplifier is capable of achieving above
95% spike detection accuracy with hundreds of nanowatt power consumption per channel.
This design indicates a promising approach to conduct spike-detection on-chip with low
power consumption and demonstrates the potential of a hybrid memristor/CMOS circuit
for power-efficient large-scale neural interfacing application.
Contents
List of Figures
List of Tables
Declaration of Authorship
Acknowledgements
1 Introduction 1
1.1 The Topic Area, Challenges and Motivation . . . . . . . . . . . . . . . . . 1
1.2 The Proposed Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Research Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Contribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 List of Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.6 Thesis Organisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2 Neural Information Processing 8
2.1 Neural Signals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Neural Electrode Techniques . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3 Neural Front-End . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3.1 Overview of Neural Front-End . . . . . . . . . . . . . . . . . . . . 13
2.3.2 Offset Rejection: AC-Coupled vs DC-Coupled Solution . . . . . . . 14
2.3.3 Neural Signal Processing: Spike Detection vs Spike Sorting . . . . 17
2.3.4 State-Of-the-Art Neural Front-Ends . . . . . . . . . . . . . . . . . 19
2.4 Main Reference Circuit: Neural Pre-amplifier . . . . . . . . . . . . . . . . 20
2.4.1 Design Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.4.2 Operational Principles . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3 Integrating Memristive Technology into CMOS Design 27
3.1 Verilog-A Memristor Model . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2 Integration of Memristor Model into Cadence . . . . . . . . . . . . . . . . 29
3.2.1 Import Verilog-A Model into Cadence . . . . . . . . . . . . . . . . 29
3.2.2 Simulation Setup for Memristor-based Circuit . . . . . . . . . . . . 30
3.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4 A Memristor-based Pre-amplifier Topology 33
4.1 Transistor Level Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
CONTENTS
4.2 Theoretical Analysis of Performance Metric . . . . . . . . . . . . . . . . . 35
4.2.1 Overview of Performance Metric . . . . . . . . . . . . . . . . . . . 35
4.2.2 Input Range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.2.3 Open Loop Voltage Gain . . . . . . . . . . . . . . . . . . . . . . . 38
4.2.4 Noise Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.2.5 Tuneable Range and Sensitivity . . . . . . . . . . . . . . . . . . . . 41
4.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 44
5.1 Transient Simulation Setup . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.1.1 Input Range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.1.2 Voltage Gain and Linearity . . . . . . . . . . . . . . . . . . . . . . 48
5.1.3 Input Offset Voltage . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.1.4 Tunable Range and Sensitivity . . . . . . . . . . . . . . . . . . . . 51
5.2 PSS Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.2.1 Open Loop Voltage Gain and Bandwidth . . . . . . . . . . . . . . 52
5.2.2 Noise Performance and Detection Accuracy . . . . . . . . . . . . . 53
5.2.3 CMRR and PSRR . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.3 Power Consumption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.4 Clock Skew and Jitter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.5 Simulation with Spike Train . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.5.1 Spike Train Generation . . . . . . . . . . . . . . . . . . . . . . . . 59
5.5.2 Spike Train Simulation . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6 A Memristor-based DC-Coupled Pre-amplifier Topology 65
6.1 The Proposed Circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.2 Transient Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
6.3 PSS Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
6.4 Tuneable Offset Range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.5 Noise Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.6 Oversampling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.7 Clock Skew and Jitter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.8 Spike Train Detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6.9 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
7 Layout Design and Chip Testing 75
7.1 Layout Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.2 Post-Layout Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.3 Strategy of Chip Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
7.3.1 Programming Memristive Device . . . . . . . . . . . . . . . . . . . 80
7.3.2 Measurement of the Bias Current . . . . . . . . . . . . . . . . . . . 81
7.3.3 Determining the Operation Phases . . . . . . . . . . . . . . . . . . 83
7.3.4 Voltage Gain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
7.3.5 Noise, CMRR and PSRR . . . . . . . . . . . . . . . . . . . . . . . 85
7.3.6 Tunable Range and Sensitivity . . . . . . . . . . . . . . . . . . . . 85
CONTENTS
7.4 Precise Signal Generation Board for Chip Test . . . . . . . . . . . . . . . 86
7.5 Pre-amplifier Chip Test . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
7.5.1 Testing on the Chip without Memristor . . . . . . . . . . . . . . . 87
7.5.2 Testing on the Chip with Memristor . . . . . . . . . . . . . . . . . 88
7.5.3 Circuitry Improvements . . . . . . . . . . . . . . . . . . . . . . . . 90
7.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
8 A Memristor-Based ∆ΣADC 92
8.1 Architecture of ∆ΣADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
8.2 Performance ∆ΣADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
8.3 Simulation with Spike Train . . . . . . . . . . . . . . . . . . . . . . . . . . 96
8.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
9 Conclusion and Future Work 98
9.1 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
9.2 Recommendation for Future Work . . . . . . . . . . . . . . . . . . . . . . 99
Bibliography 101
List of Figures
1.1 Structure comparison among the previous works in digital and analogue
forms and our proposed work. . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.1 Dendrites, cell bodies, axons, and axon terminals make up a neuron. More-
over, the transmembrane potential of the neuron can be captured by the
electrodes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Architecture and dimensions of Neuropixels 2.0. . . . . . . . . . . . . . . . 10
2.3 Architecture and dimensions of mesh electrodes. . . . . . . . . . . . . . . 11
2.4 The architecture and dimensions of multi-shank electrodes, where SU-8 is
a type of negative photoresist utilised as the core layer for waveguide. . . 12
2.5 Electrical model of the electrode between the brain tissue and CMOS de-
vice, where VDL,CDL and RDL are the double-layer potential, double-layer
capacitance and resistance, respectively, and REis the resistance from the
electrode to the CMOS device. . . . . . . . . . . . . . . . . . . . . . . . . 13
2.6 AC-coupled neural amplifier with capacitive-feedback network (CFN) topol-
ogy. The resistor Rfis a MOS-bipolar pseudo-resistor. . . . . . . . . . . . 15
2.7 The general block diagram of the chopper shows the principles and signal
transmission in the frequency domain. . . . . . . . . . . . . . . . . . . . . 15
2.8 The conventional chopper-stabilised neural amplifier. . . . . . . . . . . . . 16
2.9 General schematic of DC servo loop. . . . . . . . . . . . . . . . . . . . . . 17
2.10 One-bit delta-sigma modulator in s-domain. . . . . . . . . . . . . . . . . . 17
2.11 AC-coupled capacitive-feedback front-end. . . . . . . . . . . . . . . . . . . 19
2.12 The architecture of one-bit delta-sigma ADC where utilises three scales of
current sources to conduct calibration. . . . . . . . . . . . . . . . . . . . . 20
2.13 The architecture of pre-amplifier. It consists of three parts: (i) the core
integrating amplifier, (ii) the dynamic latch comparator and (iii) the cur-
rent bias control unit. Control signals, including clk,clk ana,clk anabar,
clk rst are all assumed to be generated by voltage sources which are strictly
periodic. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.14 Transient simulation of one detection cycle of the neural spike. One detec-
tion with four phases completes in 350ns as an example, which has been
labelled and highlighted with different backgrounds. It contains (i) reset,
(ii)integration, (iii) digitisation and (iv) off phases. (a) presents the control
of clocks; (b) compares the amplifier output (midb) with the drain of the
input transistor (drain b); (c) presents the digital output from DLC. Input
signals are set as: ina = 1V+ 50µV and inb = 1V.. . . . . . . . . . . . . 24
LIST OF FIGURES
2.15 Pre-amplifier basic functionality test: Input A (ina) is slowly swept between
[1V100µV , 1V+100µV ] over 2ms while the pre-amplifier is carrying out
a conversion every 10µs to detect the relationship between inputs A and B.
Input B (inb) remains stable at 1 Vthroughout. In this test, the amplifier
was balanced (R1 = R2). When Vina < Vinb, the left branch current is
larger than the right branch current, inducing Vmida Vmidb >0. The DLC
captures this relation and generates binary signals: Vouta = 1 and Voutb = 0,
which appears in the bottom panel as a predominantly orange output trace.
Conversely, when Vina > Vinb,Vouta = 0 and Voutb = 1, which appears as
a combined orange/blue output trace. Note: this type of simulation can
also be used to test the offset tuning range and tuning sensitivity on the
resistive state of memristive devices. When R1> R2, Vina must be lower
than Vinb to ensure a balanced output, creating an offset, this is read in the
output trace as an encroachment of the blue region into the orange (and
vice versa for R1< R2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.1 The operation sequence of importing memristor into Cadence. (a) Process
of building a new cell view in Verilog-A type for memristor. (b) Symbol
of the memristor. (c) Process of creating a symbol for memristor from
Verilog-A cell view. (d) The parameters of the memristor can be modified
in the object properties window. . . . . . . . . . . . . . . . . . . . . . . . 30
3.2 (a) Testbench of applying pulse and triangular wave to memristor to con-
duct the write and read operations in the transient simulation. (b) Pulse
chain that provokes the memristor. The input signal sequence can be di-
vided into two parts: triangular wave and pulse. For all the simulations of
our device, the read voltage is defined as Vread = 0.5Vwith 1ms duration.
The pulses can be determined with specific duration/width (tw,R), am-
plitude (Vb) and the number of pulses to provoke the memristor. With the
combination of two stages, the RS of the memristor model can be tracked
for each stimulus. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.3 Simulation results for the quadratic memristive Verilog-A model in Ca-
dence. (a) shows the responses of the model to 1,10,100µs pulse widths
with constant amplitudes (±2V). (b) presents the model response to the
voltage ramp from 1V to 3V. . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.1 The schematic of the proposed pre-amplifier. Compared to the initial design
in Figure 2.13, this circuit uses external clk to control that the DLC can
be fed by sufficient voltage difference and increase detection accuracy. . . 34
4.2 Part of schematic of core amplifier aiding the input range analysis. . . . . 37
4.3 The circuit diagram of M6 and R1 with a test voltage source presented as
(a) a large signal and (b) a small signal. Rorepresents the drain-source
resistance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.1 Block diagram of the proposed neural recording front-end. The open-loop
network OTA filters the raw data to obtain the MUA. The minute signals
charge the load capacitors (CL1,2) to realise the integration and boost the
voltage difference of mida &midb, which triggers the DLC to process
digitisation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
LIST OF FIGURES
5.2 Transient simulation of one detection cycle of the neural spike. One detec-
tion with four phases completes in 250ns, under the condition of Vina =
1V+ 50µV and Vinb = 1V. It contains (i) reset, (ii) integration, (iii) de-
tection and (iv) off phases. Full detection cycle is presented, including (a)
control signals, (b) output of the amplifier (midb) and drain voltage of in-
put transistor (drain b), (c) differential output of the OTA (∆Vmid) and
(d) digital outputs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.3 Input range results of pre-amplifier. In this simulation, the common mode
voltage was swept from zero to 1.8Vwith 0.1Vdifferential input. For
VCM [0 1.3]V, we notice that (a) Vmidb reaches sufficiently high voltage
to prompt a stable output from the (c) DLC for our chosen differential
input the output is always correct. However, the core’s analogue gain in
(b) is maximised in the narrower range [0.8,1.2]V.. . . . . . . . . . . . . 47
5.4 Intermediate differential output Vmid evolution as a function of VC M .
Differential input voltage is 50µV and the integration phase is not time-
constrained (see Figure 5.3). Voltage traces for different VCM s follow each
other closely except in the edge cases VCM {0.7V , 1.2}.. . . . . . . . . . 47
5.5 Simulation results of the differential gain analysis. In this simulation, inb
was set at 1Vwhile ina was swept from 1V500µV to 1V+ 500µV with
in steps of 10µV . (a) Vmid throughout an intentionally excessively long
integration phase. As Vmida,b increases, the cascode transistors eventually
triode causing the gain to peak and decrease. Peak gain times occur at t=
237ns and are aligned within 1ps difference. An indicative integration time
leaving a substantial margin for error can be set to, e.g. 220ns (dashed line
in (a)). (b) Output voltage difference Vmid at integration time τ= 220ns
vs input differential voltage. A linear curve excellently fits the result. The
gain is constant at approx. G= 20. . . . . . . . . . . . . . . . . . . . . . . 49
5.6 Pre-amplifier basic functionality test: Input A (ina) is slowly swept between
[1V500µV , 1V+500µV ] over 2ms while the pre-amplifier is carrying out
a conversion every 1µs to detect the relationship between inputs A and B.
Input B (inb) remains stable at 1Vthroughout. In this test, the amplifier
was balanced (R1 = R2). . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.7 Histogram of Monte Carlo simulation with 200 samples of the input offset
voltage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.8 OTA gain and input noise as a function of frequency. (a) The open-loop
network OTA presents bandpass filtering with a mid-band gain of 26 dB.
(b) Output noise of the detection cycle from (periodic) noise simulation. . 53
5.9 Principle of periodic noise jitter analysis at the threshold point. . . . . . . 54
5.10 Detection accuracy of different Vin from transient noise simulation under
the conditions of R1 = R2 = 10k (blue) and R1 = R2 = 250k (orange).
The Gaussian fittings and input-referred noise/standard derivation (σ) are
given. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.11 CMRR and PSRR of the open-loop network pre-amplifier. . . . . . . . . . 55
5.12 Clocking of the AC-coupled front-end with the jitter effect presented in dash
lines. The timing diagram presents four phases: i) reset, ii) integration, iii)
digitisation and iv) off phases. . . . . . . . . . . . . . . . . . . . . . . . . . 57
LIST OF FIGURES
5.13 Simulation results of (a) differential output voltage of OTA (∆Vmid) and (b)
output voltage (Vmid). The clk signal occurs at 220ns ideally and samples
the output of OTA, whilst the tolerant variation range is within orange
dash lines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.14 Schematic of spike train generation, where the output is at the port ‘AP’. 59
5.15 The timing diagram of spike train generation. (a) is sine wave with the
amplitude of 500µV and frequency of 10kHz. (b) is the periodic pulse,
and (c) is the generated spike train. . . . . . . . . . . . . . . . . . . . . . 60
5.16 The block diagram presents the module and connection of the testbench. . 60
5.17 The timing diagram of the control signal and spikes. (a) clk ana controls
the on and off status of the OTA. (b) Spike train (AP ) from spike generator.
(c) shows the spike with 15.77mV DC offset appended. (d) presents the
input voltages of OTA. (e) is the differential input voltage of OTA. . . . . 61
5.18 Simulation result of spike detection. (a) presents control signals which are
the same as Figure 5.2. (b) presents the differential input voltage ina inb
and the ideal spike. (c) shows the differential output voltage of OTA, and
(d) presents the digital outputs. . . . . . . . . . . . . . . . . . . . . . . . . 62
5.19 Simulation result of spike detection within 400µs. (a) presents the ideal
spike. (b) is digital output in noiseless simulation, and (c) is obtained from
transient noise simulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.20 The timing diagram is to present the spike detection accuracy. The detec-
tion cycle is 10µs in (a) and 5µs in (b), respectively. . . . . . . . . . . . 63
6.1 DC-coupled front-end. (a) Block diagram of the neural recording front-end.
(b) The electrical model of the tissue-electrode interface. (c) Simplified
schematic of the OTA assembled with memristive devices. . . . . . . . . . 66
6.2 The timing diagram of one neural detection cycle (250ns). (a) clocking
scheme. (b) input voltage of the front-end from the electrode. (c) the
output voltage of the core amplifier. (d) the differential output voltage
of the amplifier. (e) digital outputs from the DLC. (In this simulation,
ina = 1V+ 50µV and inb = 1V.) . . . . . . . . . . . . . . . . . . . . . . . 67
6.3 Detection accuracy of front-end system vs differential input voltages Vin
from 10µV to 400µV . In this case, R1 = 10kΩ, R2 {10,49,83.6,127.3}k
in order to compensate the DC offset of {1m, 10m, 50m}V.. . . . . . . . 70
6.4 Spike detection accuracy under oversampling. The initial value takes from
Fig. 6.3, where the accuracy of R2 = 83.6k and R2 = 127.2k are 67.6%
and 58.4% respectively under Vin = 50µV .. . . . . . . . . . . . . . . . . 71
6.5 Clocking of the AC-coupled front-end with the jitter effect presented in
dash line. The timing diagram presents three phases: i) reset, ii) detection
and iii) off phases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6.6 Simulation result of spike detection. (a) presents control signals which are
the same as Figure 6.2. (b) is the spike train generated in Figure 5.16.
(c) presents the input voltages of OTA. (d) shows the differential output
voltage of OTA, and (e) presents the digital outputs. . . . . . . . . . . . . 73
6.7 The timing diagram is to present the spike detection accuracy. The detec-
tion cycle is 10µs in (a) and 5µs in (b), respectively. . . . . . . . . . . . 74
LIST OF FIGURES
7.1 Differential input and cascode pair in both (a) schematic and (b) layout
view. The bulks of the above components are connected to the source of
M4&M5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
7.2 Schematic and layout of the programming circuit of the memristive device.
(a) Terminals A and B can be the user terminal and programming terminal
or in reverse. The status of memristive in the application or programming
is controlled by two signals, EN MEM and ENBAR. In layout (b), the
memristive device is in the middle and connects to PMOS (upper device)
and NMOS (lower device). . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.3 The schematic (a) and layout (b) of ESD Cell based on TSMC180nm tech-
nology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.4 Layout of the core pre-amplifier circuit with the input and output on the
boundary to connect the ESD cell or the pad directly. . . . . . . . . . . . 78
7.5 Complete layout design of the pre-amplifier. This design has two types of
pads: i) the commercial pad from the foundry and ii) the in-house pad for
programming the memristive devices. . . . . . . . . . . . . . . . . . . . . . 78
7.6 The input voltage ramp is applied to the circuit for the post-layout circuit
to determine the inherent offset voltage. The figure shows that the inherent
offset voltage is Vos =Vina Vinb = 120µV .. . . . . . . . . . . . . . . . . 79
7.7 Timing diagram for post-layout simulation. . . . . . . . . . . . . . . . . . 80
7.8 The schematic of utilising two OPA191 to generate and test precise ana-
logue signals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
7.9 Full view of the chip without the memristor electrode. . . . . . . . . . . . 88
7.10 The schematic view presents the connection inside the chip where there
exist open circuit due to missing memristor. . . . . . . . . . . . . . . . . . 88
7.11 Testing Result of the DLC with floating inputs with different pulse widths.
In this test, the pulse in the duration of 10ms with four pulse widths
{2ms, 4ms, 6ms, 8ms}are fed into the chip (presented at the top trace
of sub-figures). The digital outputs are presented in the bottom trace of
each sub-figure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
7.12 The pre-amplifier chip with the memristor pads and wiring on the top (in
white). But the wirings across the padring short the power. . . . . . . . . 90
7.13 The full layout with new wiring route for memristors and electrodes that
prevents shorting circuit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.14 The full layout with new wiring route that prevents shorting circuit. . . . 91
8.1 The block diagram of proposed ∆ΣADC. . . . . . . . . . . . . . . . . . . 92
8.2 The schematic of OTA of ∆ΣADC. (a) The OTA is the input and the
integrator interface in the ∆Σ modulator. (b) The IDAC shown on the
right creates a current feedback loop. . . . . . . . . . . . . . . . . . . . . . 93
8.3 The timing diagram of ∆ΣADC. (a) the input of the system. (b) the output
of the OTA. The one-bit IDAC provides the modulation. (c) and (d) are
the outputs of the DLC, respectively. . . . . . . . . . . . . . . . . . . . . . 95
8.4 Performance of ∆ΣADC. (a) the output spectrum for the input sine wave
of 10 mVpp under the noise simulation. (b) SNR versus input level where
indicates 55dB DR. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
LIST OF FIGURES
8.5 Left: the timing diagram of ∆ΣADC with (a) the spike train input with
15.77mV DC offset. (b) the output of the OTA. (c) and (d) are the outputs
of the DLC, respectively. Right: output spectrum for spike train with the
maximum amplitude of 500µV and 15.77mV DC offset. . . . . . . . . . . 97
List of Tables
2.1 Summary of the characteristics of different neural signal modalities. . . . 10
2.2 Comparisons of the novel electrodes. . . . . . . . . . . . . . . . . . . . . . 12
2.3 Parameters of the electrical model from the measurement of different ma-
terials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4 Comparison among state-of-art neural front-end and the specification of
this work. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.1 Specification of the memristive integrating OTA. . . . . . . . . . . . . . . 36
5.1 Specification and overview of simulation methodology for the memristor-
based OTA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.2 Results of a normal distribution from Monte Carlo simulation. . . . . . . 50
5.3 The offset voltage of pre-amplifier vs memristor device resistive state is
quoted at 1µV resolution. . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.4 Performance metrics of the discrete-mode threshold detection system. . . 56
6.1 Sizes of devices in the proposed architecture, where the bias current of the
core amplifier is Itail = 3µA.R3 is replaced by a diode-connected NMOS.
The detailed schematic is in Figure 4.1. . . . . . . . . . . . . . . . . . . . 65
6.2 Performance metric of memristor amplifier (The energy and average power
consumption are of one detection cycle with 250ns, including both mem-
ristor amplifier and DLC). . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.3 The compensated DC offset versus the resistive state of memristors under
R1 = 10 kΩ. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.4 Memristor tuneable offset range when R1 = 10k and R2 is from 10k to
130kΩ. The third column presents the tuneable range with 1% RS variation
from two memristors. The offset resolution is 10µV .. . . . . . . . . . . . 69
7.1 Sizes of devices in the proposed architecture, where the bias current of core
amplifier is Itail = 3µA.R3 is replaced by a diode connected NMOS. . . . 76
7.2 The pins of the pre-amplifier with the classification of type. There are
some abbreviations: ana in/out represents analogue input/output signal,
ana in HV means high voltage analogue input which is 5Vwith the sup-
ply of VDD MEM. Besides, dig in/out is the digital input/output signal.
pwr x, where x= 1.8/5 is the supply voltage. . . . . . . . . . . . . . . . . 78
7.3 Performance metrics comparison between pre-layout and post-layout sim-
ulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
7.4 The pin connection for programming memristive devices. . . . . . . . . . . 82
7.5 The pin connection for measuring branch current. . . . . . . . . . . . . . . 83
7.6 The pin connection for measuring the reset phase. . . . . . . . . . . . . . 83
LIST OF TABLES
7.7 The pin connection for measuring the integrating phase. . . . . . . . . . . 84
7.8 The pin connection for measuring the DLC only. . . . . . . . . . . . . . . 89
8.1 Sizes of devices of OTA in the proposed ∆Σ ADC. . . . . . . . . . . . . . 94
8.2 The compensated DC offset versus the resistive state of memristors under
the condition of R1 = 1 kΩ. . . . . . . . . . . . . . . . . . . . . . . . . . . 94
8.3 Performance of the proposed ∆ΣADC under the condition of transient noise. 96
8.4 The critical transistor dimensions and noise contribution of the OTA. . . 96
Declaration of Authorship
I declare that this thesis and the work presented in it is my own and has been generated
by me as the result of my own original research.
I confirm that:
1. This work was done wholly or mainly while in candidature for a research degree at
this University;
2. Where any part of this thesis has previously been submitted for a degree or any
other qualification at this University or any other institution, this has been clearly
stated;
3. Where I have consulted the published work of others, this is always clearly at-
tributed;
4. Where I have quoted from the work of others, the source is always give. With the
exception of such quotations, this thesis is entirely my own work;
5. I have acknowledged all main sources of help;
6. Where the thesis is based on work done by myself jointly with others, I have made
clear exactly what was done by others and what I have contributed myself;
7. Parts of this work have been published as journal papers and conference contribu-
tions in the list of publications.
Signed:.......................................................................... Date:..................
]
Acknowledgements
I am grateful to those who have accompanied me on the path to my PhD.
My supervisors, Professor Themis Prodromakis and Alexantrou Serb, provided me with
the invaluable opportunity to pursue a PhD. I would like to express my sincere gratitude
to them. Over the years, I have appreciated your consistent guidance, encouragement,
and patience. I would also like to thank my internal examiners, Professors Geoff Merrett
and Koushik Maharatna, for their insightful and helpful comments on my work.
I would express appreciation to my colleagues at the University of Southampton: Shiwei
Wang, Sachin Maheshwari, Spyros Stathopoulos, Yihan Pan and Patrick Foster. It has
been a pleasure collaborating with you. I cherish the time we collaborated, brainstormed,
discussed and shared our ideas, working on a single project. And I would like to thank
my colleagues at Imperial College London: Christos Papavassiliou, Peilong Feng and Lijie
Xie, who provided me with excellent assistance and convenience during my chip testing
at Imperial College London.
I would like to express my heartfelt gratitude to my grandparents, parents, aunts and
cousin. Their enduring affection and encouragement have been my primary source of
support during my years abroad. I would like to express my most profound appreciation
to my friends Franco and Vivian, as well as my cats Tyson and Oscar. Thanks for keep
accompany with me.
Lastly, I express my gratitude towards The Royal Society for funding the industry fellow
PhD student scholarship.
Chapter 1
Introduction
1.1 The Topic Area, Challenges and Motivation
Recording and monitoring neural signals using implantable microsystems is essential to
understanding biological activity and functions. These can be utilised to advance diagnos-
tic and therapeutic solutions [1] and overall neuroscience research [2]. From an application
point of view, the ability to record and process extracellular neural signals can lead to
novel therapeutic neuromodulation solutions for treating disease [3] and new opportuni-
ties in prosthesis [4]. For these purposes, high-density micro-electrode arrays (MEA) and
recording systems are required to acquire and process a large number of neuronal signals
simultaneously in real-time.
With the development of techniques in neuroscience, the micro-electrode arrays can access
a more significant number of neurons chronically in the order of 10,000 [5;6]. Scientists
are developing higher-density implantable recording systems [7;8;9]. Besides, there is a
strong commercial interest in such approaches, with Neuralink [10] proposing a package
with 3072 electrodes and a small recording system that delivers amplification and digi-
tisation capabilities. However, we are supposed to overcome several technical challenges
to construct an extracellular neural signal recording with a high density of up to 10,000
channels. To give a better perspective, the prior arts, followed by the challenges and
motivation of this research field given below.
Neural signal recording can be realised through the approaches shown in Figure 1.1 (in
black). Such signals are transduced by a physical interface consisting of probes and elec-
trodes [11;12]. These raw neural signals generally range around 50500µV in amplitude
with 5 10µV background noise added on top and span around 0.1Hz to 10kH z in fre-
quency [13]. The raw signals are passed to a front-end, which amplifies the signal to the
milli-volt range. The amplified waveform is then filtered in a filter module, removing in-
terference (e.g. power line interference) whilst passing Action Potentials (AP). After that,
the signal is passed into the back-end, which refines the signal and generally attempts to
1
2Chapter 1 Introduction
extract the computationally important information. One approach is to process the signal
via analogue processing methods [9]. Alternatively, it can be digitised by an ADC and
subsequently post-processed in the digital domain [7;8]. After that, the neural signal can
be transmitted off-chip for further processing and/or storage.
Proposed front-end pre-amplifier
Detection of neural spikes and digitalization
Electrode Amp Digital Signal
Processing
DigitizerFilter
Analogue Signal
Processing
Figure 1.1: Structure comparison among the previous works in digital and analogue
forms and our proposed work.
The state-of-the-art in-vivo systems consist of two schemes: ‘high-channel count but wired’
[7;8;9] and ‘wireless but low-channel-count’ recording system [14;15;16]. As for the wired
system, the front-end channels are on the implanted chip. With tethered cables connected,
the data processing module for high-channel-count is on the external device, which is able
to process further computation. To the best of our knowledge, Neuralink [10] published
a wired recording system that can achieve 3072 recording channels. However, the wire
connection not only limits the mobility of observed animals, but also induces the risk of
infection and device breakage [17]. The alternative solution, a fully wireless system, can
overcome the above challenges. However, the high power/area budget of data transmission
constrain scaling up the system. As for the wireless system, the number of recording
channels is limited to around 100 [15]. It is because the transmission module occupies
20%50% of the power budget depending on the real-time transmission speed. To reduce
power consumption, scientists focus on these three approaches: resource sharing, power
scheduling and supply voltage reduction [3]. However, following the conventional path will
limit them to enhance an order of magnitude. In this research, we target both wireless
and high-channel-count recording systems.
The systems are required to be low-power to prevent damaging the surrounding tissues and
extend the battery life [18]. The international guideline stated that the power consumption
of the implantable neural interface should be confined within 1540mW depending on the
depth that the interface is implanted [19]. Chen et al. [20] have studied the state-of-the-art
wireless neural systems and summarised that the transmission module occupies 20%50%
of total power consumption depending on the volume of signals to be transmitted and
Chapter 1 Introduction 3
the system functionalities. The power constraint poses critical challenges to developing
high-channel-count recording solutions, limiting the monitored neurons in the order of
100-1000 simultaneously [21].
As mentioned above, we set the maximum power consumption of the interface to 15mW .
To improve the power efficiency, we proposed a hybrid Complementary Metal–Oxide–
Semiconductor (CMOS)/Memristor design that directly detects and classifies neural sig-
nals. It bypasses several steps in the conventional approach that brings in the possibility of
building both wireless and high-channel-count neural interfaces (Figure 1.1). It is because
we do not need to record the complete waveform or with high resolution of digitisation
by taking advantage of the memristor-based neural signal processing, which is evidenced
in the prior work [22]. Only 1-bit digital signals need to be transmitted off-chip; however,
we target 10,000-channel data. Therefore, we estimated that data transmission account
for 50% of the power budget which is 7.5mW . Additionally, Fassio et al. [23] published
an ultra-low-power biasing circuit that consumes picowatt-level power; thus, we omitted
the power consideration of the biasing circuit temporarily. Therefore, the rest modules
of our system consume 7.5mW . In addition, we target a 10,000-channel system which
indicates we need to constrain the power consumption of each channel to 750nW . The
detailed solution is given in the following section.
1.2 The Proposed Solution
This research presents an alternative paradigm by developing a front-end pre-amplifier
concept that can operate in a start-stop scheme to achieve nanowatt-level power consump-
tion per channel. This processing scheme conducts the amplification from microvolts to
millivolts, spike detection and digitisation at the first stage (Figure 1.1 in orange). Thus,
the requirement on afterwards modules is relaxed to processing digital signals represent-
ing neural spikes rather than the entire signal. This approach led to a design concept
translating a train of analogue signals into digital, reducing the critical data points from
the first stage.
The proposed front-end pre-amplifier consists of an integrating amplifier and a Dynamic
Latch Comparator (DLC), conducting amplification and digitisation. As opposed to tra-
ditional practice where an ‘expensive’ amplifier is used to deliver high gain and low noise,
here, we opt to integrate the microvolt-level signal to millivolts. The integration process
not only filters out noise but also provides amplification; thus, we refer to this as an
integrating amplifier. This fully differential system approach allows for integrating the
difference between the input neural signal and a reference to a voltage level which triggers
a threshold detector. Three operations (amplification, filtering and digitisation) can thus
be realised within a single module. Further back-end processing is possible on the data
emitted by our system in principle.
4Chapter 1 Introduction
One of the most significant features is that the memristor is introduced as a trimming de-
vice to compensate for the offset voltage. Memristors are metal-oxide-metal devices whose
Resistive States (RS) can be controlled by applying appropriate bias voltages. Memristor
has been proven to be compatible with CMOS technique in terms of processes and mate-
rials [24;25]. Memristor is constructed as a stack by the bottom electrode, active layer
and top electrode, where both top and bottom electrodes can connect with peripheral
and control CMOS circuitry, e.g. the drain/source of the transistor. In addition, mem-
ristors can be directly fabricated on the top of the CMOS circuits (above passivation)
and connect to the transistor through Tungsten via, which increases area efficiency [26].
Being protected by passivation, the CMOS circuit will not be affected by both fabrication
processes and materials of the memristor. We utilised the in-house P t/Al2O3/T iO2/P t
memristor in our designs since it can achieve 92 distinct resistive states from 20k and
120kΩ, which can be programming in the precision of 1.09k [27]. Thus, we can tune the
offset with the precision of the microvolt level by operating the memristor as a trimming
device.
In addition, it can be used in various applications ranging from non-volatile memory
[28], neuromorphic networks [29] to programmable resistances [30]. Its non-volatility,
nanometer scale and CMOS-compatible processing [31;32] add to the benefits of using
memristors in our system as trimming devices. Such a passive trimming element adds no
power dissipation in principle except when the device needs to be programmed to tune the
circuit. Its high non-volatility features the architecture low maintain power consumption,
compared with Dynamic Random-Access Memory (DRAM) approach that needs to be
refreshed to preserve the data [33]. Besides, compared with the flash memory whose
program and erase voltages are typical 1521V[34], the low programming voltage (below
3Vin [27]) simplifies the actual chip design. Incidentally, memristors have also been
demonstrated to discriminate spiking from background activity [22] when used directly as
sensing elements (alas on an already amplified waveform). It is a distinct approach that
nicely complements this programme of research.
Throughout this work, memristors are operated under a ‘write once read many’ regime
where they are used to trim differential current paths and, as a result, tune the threshold
(for spike detection) voltage of our system. Moreover, to program the device, a pair of
single-transistor switches connecting the devices to an appropriate programming voltage
and a trans-impedance amplifier (all shared across multiple channels) can perform the
programming. In contrast, the main amplifier being programmed is shut down. Our
research focuses exclusively on the function of the channel in regular operation, where the
memristor resistance does not change. Therefore, this specific programming regime will
not issue energy efficiency or complexity in our design.
Chapter 1 Introduction 5
1.3 Research Objectives
For this research program, we aim to construct a hybrid CMOS/Memristor neural interface
that can be used for the real-time processing of neural signals. To achieve the aims of the
research, the following objectives are identified:
1. Develop and validate power and area-efficient CMOS/Memristor neural interface
circuitry that can directly detect and classify neural signals.
2. Fabricate and demonstrate that the developed circuitry (single channel) can achieve
sub-microwatts average power consumption with an implementation in silicon.
3. Prove the proposed approach for large-scale neural interfacing application in the
experiment and demonstrate that the developed chip can readout and process neural
signals.
1.4 Contribution
The novel contributions presented in this dissertation are listed below.
1. Made the memristive Verilog-A quadratic model available to Cadence [35].
Previously, our group only had the exponential Verilog-A model under the 10 17k RS
range, which limits the utilisation of memristor in circuit design. In [35], I converted the
mathematical model from [36] into the Verilog-A model which owns a wider RS range
from 20k to 120kΩ. The methodology of building the Verilog-A model refers to [37].
The Verilog-A model and methodology of integrating memristor into Cadence have been
included in Chapter 3.
2. Established methodology for evaluating performance metric of dynamic
pre-amplifier [38;39]. The pre-amplifier implements the amplification by integrating
load capacitors. Thus, it features a dynamic operating point, rendering traditional anal-
ysis (AC/DC) unsuitable. In this work, I analysed the transistor-level pre-amplifier and
developed the methodology to complete the performance metric by combining transient
simulation and periodic steady-state analysis. This benchmarking approach is finally
leveraged for providing useful insights and design trade-offs of the memristor-based inte-
grating amplifier in Chapter 4and Chapter 5.
3. Studied the impact of memristor IV non-linearity on the pre-amplifier
[40]. In this design, I calibrated RS measurements to the operating tail current. The
measurement illustrates that the small deviations from tail current during integration do
not materially change the extreme offset trimming precision or the overall performance.
This was the last conceptual bottleneck identified before practical implementation, which
has now been overcome. It can be accessed in Chapter 4.
6Chapter 1 Introduction
4. Designed memristor-based DC-coupled pre-amplifier [41]. Instead of using
huge capacitors at the input to reject the DC offset (AC-coupled solution), in this work,
we utilised the memristive device to tune the offset in the wider range that can cover up
to 50mV (DC-coupled solution). The full performance metric is elaborated in Chapter 6.
The simulation results present that the DC-coupled pre-amplifier can achieve 95% detec-
tion accuracy while the power consumption is down to 120nW . Circuit design and
simulation details are given in Chapter 6
5. Designed memristor-based first-order ∆ΣADC. The pre-amplifier operating as
a conventional static amplifier can also be utilised in the ∆ΣADC, which rejects up to
50mV DC offset by calibrating memristors. The designed transistor circuit can achieve
8.51 ENOB under the transient noise in the range of [1Hz, 100M H z] (in Chapter 8).
1.5 List of Publications
Journal Articles
A. Serb, I. Kobyzev, J Wang and T. Prodromakis (2020) “A semi-holographic hyperdi-
mensional representation system for hardware-friendly cognitive computing” Philosophical
Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences,
378 (2164), 1-18. (doi:10.1098/rsta.2019.0162).
J. Wang, A. Serb, C. Papavassiliou, S. Maheshwari, and T. Prodromakis, (2021) “Analysing
and measuring the performance of memristive integrating amplifiers” International Jour-
nal of Circuit Theory and Applications, 49(11), 3507-3525
S. Maheshwari, S. Stathopoulos, J. Wang, A. Serb, Y. Pan, A. Mifsud, L.B. Leene, J.
Shen, C. Papavassiliou, T.G. Constandinou, and T. Prodromakis, (2021) “Design Flow
for Hybrid CMOS/Memristor Systems—Part I: Modeling and Verification Steps” IEEE
Transactions on Circuits and Systems I: Regular Papers, 68(12), 4862-4875
S. Maheshwari, S. Stathopoulos, J. Wang, A. Serb, Y. Pan, A. Mifsud, L.B. Leene, J.
Shen, C. Papavassiliou, T.G. Constandinou, and T. Prodromakis, (2021) “Design Flow
for Hybrid CMOS/Memristor Systems—Part II: Circuit Schematics and Layout” IEEE
Transactions on Circuits and Systems I: Regular Papers, 68(12), 4876-4888.
Conference Proceedings
J. Wang, A. Serb, C. Papavassiliou, and T. Prodromakis (2021) “Accounting for Mem-
ristor IV Non-linearity in Low Power Memristive Amplifiers”, 2021 IEEE International
Symposium on Circuits and Systems (ISCAS ) (pp. 1-5). IEEE.
J. Wang, A. Serb, S. Wang, T. Prodromakis, (2022) “Hybrid CMOS/Memristor Front-
End for Multiunit Activity Processing” accepted by The IEEE International Symposium
Chapter 1 Introduction 7
on Circuits and Systems (ISCAS ).
J. Wang, A. Serb, S. Wang, T. Prodromakis, (2022) “Offset Rejection in a DC-Coupled
Hybrid CMOS/Memristor Neural Front-End” accepted by The IEEE International Sym-
posium on Circuits and Systems (ISCAS ).
1.6 Thesis Organisation
The thesis is organised as follows. Chapter 2presents literature on neural information
processing. The introduction of applied memristive devices in terms of the operational
behaviours and methodology of integration into Cadence presents in Chapter 3. Design
concepts, the basic operation and transistor-level analysis of the pre-amplifier have been
introduced in Chapter 4. The methodology for obtaining the full performance metric of
the AC-coupled memristor-based pre-amplifier is given in Chapter 5, followed by the eval-
uation and design consideration. The pre-amplifier is also utilised to reject DC offset up
to 50mV in the form of the open-loop circuit (in Chapter 6) and ∆ΣADC (in Chapter 8).
The layout design of the single channel pre-amplifier and the testing result of the chip are
included in Chapter 7. Finally, Chapter 9contains the conclusion and the recommended
future work.
Chapter 2
Neural Information Processing
The neural recording techniques allow scientists to investigate brain activity in depth and
offer the potential to treat neuron disorders. The neural front-end should be designed after
a thorough review of the signals of interest and the techniques employed by the various
modules of the monitoring system. This chapter provides an overview of (i) characteristics
of neural signals, (ii) electrode techniques related to signal acquisition, (iii) the state-of-
the-art front-end designs (in the last five years) which feature a variety of offset rejection
and signal processing techniques, and (iv) the primary reference design for the thesis.
2.1 Neural Signals
Neurons in the nervous system communicate with other nerve cells by passing electrical
and chemical signals. A neuron consists of dendrites, cell body (soma), axon and axon
terminals (Figure 2.1). Dendrites receive signals from other neurons in a neuron cell, while
axon terminals send out messages. After dendrites are excited (by chemical signals), the
cell body transmits information to the axon as an action potential [42]. Action potential
transmission generates transmembrane potential, which the intracellular and extracellular
neural recording can capture. By placing electrodes in the region of interest, intracellular
and extracellular action potentials can be recorded [43]. Consequently, the action potential
can be recorded as the potential difference between the electrode’s tip and ground electrode
[44].
The intracellular neural recording is a highly invasive technique that the transmembrane
potential is measured by penetrating the membrane with the electrodes [45]. This tech-
nique is capable of capturing highly accurate information about the electrophysiology (e.g.
a few ion channels) and obtaining a high amplitude of the intracellular action potential
that is in the range of 70mVpp [46]. However, due to the invasion of electrodes, this tech-
nique will cause irreversible damage to the neurons, so it can only be used for fundamental
8
Chapter 2 Neural Information Processing 9
Cell Body
Dendrite
Axon
Axon Terminal
Figure 2.1: Dendrites, cell bodies, axons, and axon terminals make up a neuron. More-
over, the transmembrane potential of the neuron can be captured by the electrodes.
discoveries instead of medical interventions. As for the application, it is primarily applied
for in-vitro measurement, while it is not appropriate for long-term or chronic recording
[47].
In contrast, electrodes can be placed in the cortical tissue to record the extracellular
action potential. This technique allows chronic monitoring of large groups of neural
activities. However, compared to the extracellular recording, the recorded intracellular
action potential has a weaker amplitude, ranging from 50µVpp to 500µVpp depending on
the distance between the electrodes and the active neuron [48].
This thesis focuses on extracellular neuron recordings that are more appropriate for
chronic monitoring. The recorded signals from the electrodes comprise extracellular ac-
tion potential and Local Field Potential (LFP). The local field potential is generated by
a group of neurons around the tip of the electrodes (50 - 350 µm) [49]. In addition, when
the neuron is excited asynchronously, the action potential may be cancelled and the net
field potential will be eliminated. Therefore, the LFP can be utilised to reflect the co-
ordination of the population of neurons and the neural network dynamics. However, the
contribution of individual spiking neurons to LFP has yet to be adequately described [45].
The LFP has a frequency of between 1mHz and 200Hz with an amplitude of 500µVpp -
5mVpp.
As for the action potential, it contains both single-unit activity and multi-unit activity.
The single-unit extracellular AP owns a frequency range from 100Hz to 10kHz. As for the
multi-unit spiking, Choi et al. [50] obtained multi-unit activity by utilising a bandpass
filter from 300Hz to 3000Hz. On the other hand, Stark et al. [51] acquired the multi-
unit spiking from the root mean square of the signal in the frequency range of 300Hz to
6000Hz. In addition, Stolerman [52] only provided the vague frequency range of multi-unit
spiking from 400Hz to a few kilohertz. Since the application of this design is to predict
movement and epileptic seizure by counting the action potential that includes both single-
10 Chapter 2 Neural Information Processing
and multiple-unit spiking, we target the signal of interest in the frequency from 100Hz to
10kHz. The attributes of different neural signals are summarised in Table 2.1.
Table 2.1: Summary of the characteristics of different neural signal modalities.
Modality Amplitude Bandwidth Potential for chronic recording
Intracellular action potential 10-70mVpp 100-10kHz low
Extracellular action potential 50-500µVpp 100-10kHz high
Local field potential 0.5-5mVpp 1m-200Hz high
2.2 Neural Electrode Techniques
Neural electrodes have significantly advanced the field of neuroscience by extracting and
detecting neural signals. In addition, the extracellular action potential and local field
potential are the main focus of this thesis. Spatial integration, stability and functionality
are the three development axes for electrodes [53]. Consequently, we split the evaluation of
innovative electrode approaches into three categories. Instead of describing each electrode
in detail, we briefly describe each direction with a single example.
Spatial integration: interpreting the coordination of the brain demands a considerable
increase in the number of monitored neurons. Moreover, this requires large arrays with
multiplexed recording sites and minimum numbers of footprints [53]. The development
of complementary metal–oxide–semiconductor (CMOS) technology features the probes
with more compact multiplexing circuits (in shank) and I/O connections (in base) [54;
55]. In the thesis, we take the ‘Neuropixels’ as an example. The neuropixels 2.0 probe
[56] contains four shanks and a base presented in Figure 2.2 with the dimensions. On
the shank, 1280 electrodes are assembled, with each electrode containing an analogue
15𝜇m
70𝜇m
12𝜇 12𝜇m
250𝜇m
Figure 2.2: Architecture and dimensions of Neuropixels 2.0.
Chapter 2 Neural Information Processing 11
switch and 1-bit memory. The analogue switches determine which set of 384 electrodes
is recorded concurrently. As for the non-implantable component, the base is fabricated
by 130nm CMOS technology that comprises the electrical components responsible for
multiplexing, amplification, digitisation, and power management units. Pre-processing the
neural signals on the probe base prevents signal degradation during the data transmission
from the probe to external devices. Neuropixels probe has been utilised to track the visual
response in rats for up to two months.
Long-term stability: Access to time-dependent brain functions at the level of single
neuron and the development of cognitive processes demands an implantable probe with
long-term stability [57]. The interfaces should be featured with minimum perturbation to
the brain tissues to achieve stable chronic monitoring. Therefore, scientists have concen-
trated on developing probes with significantly reduced mechanical stiffness than silicon or
micro-wire probes [58;59;60]. To realise long-term recording, researchers (i) addressed
the distance between the probe and brain tissues and (ii) designed the probes that ’ap-
pear’ and ’behave’ similarly to the tissues [53]. Mesh electronic probes were introduced
to have similar maximum characteristic dimensions to individual neuronal somatic cells,
with the same bending stiffness values as brain tissue (Figure 2.3). It has been applied
for eight-month tracking of the age-related functional change of mouse neurons [60].
2𝜇m-width
interconnector
20-μm-diameter
recording site
Figure 2.3: Architecture and dimensions of mesh electrodes.
Multiple functionalities: compared to the above electrodes, the multi-functional can
achieve both recording the neural signals and stimulating the brain region. Multi-functional
electrodes are mainly utilised to treat neurological disorders like Parkinson’s disease and
depression by real-time controlling and correcting the disorders [61]. For example, Shin
et al. [62] proposed a multi-shank Micro-electromechanical systems (MEMS) electrode
that consists of (i) recording sites, (ii) microfluidic channels for chemical delivery, and
(iii) optical waveguides for optical stimulation in the brain. This functional integration
is essential for accurately modulating neural circuits in-vivo. The architecture of the dis-
cussed electrode presents in Figure 2.4. The comparisons of the key characteristics of the
12 Chapter 2 Neural Information Processing
electrodes in three categories are presented in Table 2.2.
SU-8 waveguide
Signal line
Microfluidic channel
20𝜇 20𝜇m
recording site
128𝜇m30𝜇m
Figure 2.4: The architecture and dimensions of multi-shank electrodes, where SU-8 is
a type of negative photoresist utilised as the core layer for waveguide.
Table 2.2: Comparisons of the novel electrodes.
Technology Multiplexity Stability
Neuropixels [56] 384 recording channels data collection from the same
neuron for more than two
months
Mesh electrodes
[60]
16-128 recording channels tracking of the same neuron
over eight months
multi-shank neu-
ral electrodes [62]
32 recording electrodes, 2-
inlets microfludic mixer and
SU-8 waveguides
causes litter immune re-
sponses over two weeks
After studying different categories of electrodes, we introduce the electrical model of the
electrode-tissue interface that translates the neural activities into electrical signals [63].
The structure of the electrical model is in Figure 2.5 and the example data is shown in
Table 2.3.
Table 2.3: Parameters of the electrical model from the measurement of different mate-
rials.
Parameter P t Electrode IrOxElectrode
VDL 15.77mV 3.12mV
CDL 200pF 1.2nF
RDL 10M100M
RE5k5k
Chapter 2 Neural Information Processing 13
VDL
CDL
RDL
RE
Neural
activities
Electrical
signals
Tissue
interface
Electrode
interface
CMOS
interface
TissueElectrochemical
double-layer
Electrode CMOS
Figure 2.5: Electrical model of the electrode between the brain tissue and CMOS device,
where VDL,CD L and RDL are the double-layer potential, double-layer capacitance and
resistance, respectively, and REis the resistance from the electrode to the CMOS device.
2.3 Neural Front-End
2.3.1 Overview of Neural Front-End
Neural signals from biology are transduced by a physical interface consisting of probes
and electrodes. Typically, the amplitude of these raw neural signals ranges from 50 to
500µV , with 5 to 10µV added background noise, and the frequency ranges from 0.1Hz
to 10kHz [13]. The neural signal recording is typically realised through the approaches
shown in Figure 1.1. Raw signals are passed to a front-end, which amplifies the signal
to the millivolt range. The amplified waveform is then filtered band-pass filter, which
eliminates interference (e.g. power line interference) while allowing action potentials to
pass. After that, one approach is to pass the signal into back-end processing in the form of
an analogue signal [64]. Alternatively, and indeed more typically, it can also be digitised
by an ADC and subsequently post-processed in the digital domain [65]. The above belongs
to front-end processing, which is presented in Figure 1.1 (black block diagram).
The signal is then transmitted to the back-end, which refines the signal and generally
attempts to extract the computationally important information. As a rule, the back-
end process consists of either spike detection or spike sorting. Spike detection involves
determining whether a neural action potential (spike) occurs at any given time. The
more complicated spike sorting process is then to also group detected spikes into clusters
depending on the similarity of their shapes, i.e. to identify groups of spikes that are likely
to originate from different neurons [66]. After processing in the back-end module, the
neural signal will usually be transmitted off-chip for further processing and/or storage.
In digital processing systems, data integrity and power consumption increase as the ADC
resolution is improved. Nevertheless, this approach has proved more prevalent and efficient
than the alternative of processing the waveform in the analogue domain. The reason is
primarily the ease of working with clean, digital signals when extracting information in
14 Chapter 2 Neural Information Processing
the back-end. Once the relevant information has been extracted, the digitally compressed
neural signal waveform is significantly lower than the original analogue waveform.
Before reviewing the state-of-art neural front-end designs, we study the scheme of DC-
offset rejection (AC and DC solutions) and the neural signal processing (spike detection
and spike sorting) in the following.
2.3.2 Offset Rejection: AC-Coupled vs DC-Coupled Solution
One of the challenges of designing a neural front-end is to reject the DC offset from the
electrodes. DC offset exists on the front-end’s input due to the electrochemical effects at
the electrode-tissue interface [67]. The electrical model is presented in Figure 2.5. The
differential DC input is 110mV typically and can reach up to 50mV . Such DC offset may
saturate the differential amplifier. Various techniques have been proposed to overcome
the input DC offset, including AC-coupled and DC-coupled solutions [68;69;70]. In the
following, we study (i) the AC-coupled solution - closed-loop capacitive feedback network,
(ii) chopper stabilisation, (iii) DC-servo loop and (iv) the DC-coupled solution by inducing
Delta-Sigma ADC in the front-end.
Closed-loop capacitive-feedback network (CFN): This architecture in Figure 2.6
is firstly introduced by Harrison [71] and widely utilised in neural amplifier subsequently
[72;73]. In this topology, the resistor Rfis realised in MOS or bipolar element. From
the frequency aspect, the large input capacitor Ciis utilised at the input to block the DC
offset from tissue. The parallel design of capacitor Cfand resistor Rfin the feedback
circuit operates as a high-pass pole: 1
RfCf. The OTA determines the low-pass pole, which
falls in the range of 1Hz. The huge capacitor of Ciis required to attain such a low cutoff
frequency. The mid-band gain of this amplifier is Ci
Cf. It leads to a large area, low
mid-band gain and low input impedance of the amplifier.
In summary, the input capacitors (in the order of picofarad) occupy a sizeable on-chip
channel area (in hundreds of µm2) and prevent scaling up. Besides, the highly resistive
triode-biased MOS transistor induces nonlinear behaviour in the presence of a large output
voltage swing and DC operating point drift due to transistor leakage. This non-linearity
causes distortion, process, voltage and temperature (PVT) variations which render the
high-pass pole frequency time-variant.
Chopper-stabilised amplifier. The chopper is the continuous-time modulation tech-
nique in which signal and offset can be modulated at different frequencies [74;75;76].
In this review, we go through the principle of the chopper (in Figure 2.7) before induc-
ing the chopper-stabilised ac-coupled amplifier (in Figure 2.8). The input voltage Vin is
transformed to a square wave voltage at the frequency fch by passing through a chopper
(see Vin). After initial signal modulation, the noise and offset are added to the modulated
signal (see V1). The modulated signal is then amplified with an input offset of its own.
Chapter 2 Neural Information Processing 15
OTA
Ci
Ci
Rf
Rf
Cl
Vip
Vin
Von
Vop
Cf
Cf
+
+
_
_
Figure 2.6: AC-coupled neural amplifier with capacitive-feedback network (CFN) topol-
ogy. The resistor Rfis a MOS-bipolar pseudo-resistor.
After amplification and the second chopper, the modulated signal is demodulated back
to DC. In contrast, the low-frequency noise and offset are modulated to the harmonics of
the rotor frequency, resulting in a chopper ripple at the amplifier output (see V2). A LPF
is then applied to remove offset and 1/f noise from the modulated low-frequency signal
(see Vout).
Vin
Low-pass
filter
Chopper
modulator
Chopper
demodulator
Vout
t
m1(t)
t
m2(t)
Vos+VN
fch fch
f/fch
Vin
1 3 5
signal
f/fch
V1
1 3 5
modulated signal
offset & noise
f/fch
V2
1 3 5
signal
modulated
offset & noise
f/fch
Vout
1 3 5
signal
V1V2
A( f )
Figure 2.7: The general block diagram of the chopper shows the principles and signal
transmission in the frequency domain.
On the other hand, the chopper often comes with the ac-coupled amplifier in the neural
front-end that is referred to as a chopper-stabilised amplifier (in Figure 2.8). Compared
to a pure chopper that employs a low-pass filter, the chopper-stabilised amplifier feeds the
demodulated output back to the input. The amplifier then eliminates the low-frequency
errors. Compared to the closed-loop capacitive-feedback network, the chopper-stabilised
amplifier is capable of achieving considerable noise and offset reduction. Nonetheless,
16 Chapter 2 Neural Information Processing
this architecture still utilises extra large input capacitors and requires extra techniques to
boost the input impedance.
Figure 2.8: The conventional chopper-stabilised neural amplifier.
DC servo loop (DSL): The general schematic of the DC servo loop in Figure 2.9
illustrates the principle of DC offset cancellation. This system is capable of measuring
the output voltage Voand comparing it to the intended DC value VREF . The voltage
difference is then sent back to the input through an integrator, yielding a DC output
voltage equal to the desired DC level (VREF ). The transfer function is
Vo(s) = Vi(s)s
1 + s/Gm +VREF (s)1
1 + s/Gm (2.1)
This architecture is able to eliminate DC components at the input by incorporating a
high-pass filter for Viand an integrator with a substantial time constant [77]. And it
functions as a lowpass filter for the VREF , which establishes the DC output level. The
conventional DC-Coupled technique is to induce a DC servo feedback loop in order to
eliminate the DC offset [70]. Pham et al. [70] lowers the large input capacitance by
inducing additional capacitors, resistors, OTAs and chopper stabilisation modules into
the DC servo loop. However, the opamp in the feedback circuit significantly increases the
power consumption, and its various open-loop gain results in a variable high-pass pole.
Even though it is effective at removing DC offset, it can be only applicable to systems
with a restricted number of channels. This is because such a system requires substantial
passive components to generate a low-frequency high-pass pole.
Delta-Sigma modulator (DSM): The delta-sigma modulator has become the main-
stream front-end for achieving low noise level [78]. The block diagram of the one-bit
delta-sigma modulator is in Figure 2.10. The analogue signal at the input will be sampled
by multiple times, which is known as oversampling. Since the sampling rate is multiple
times faster than the digital output, the individual sample is accumulated over time and
then averaged with other input samples. The feedback loop in the modulator operates
different functions to desired signal (signal transfer function - STF) and the noise (noise
Chapter 2 Neural Information Processing 17
+_
+_
ViVo
VREF
Gm
1
s
Figure 2.9: General schematic of DC servo loop.
transfer function - NTF):
Vo=Vi
1
s+ 1 +Ns
s+ 1 (2.2)
It can be obtained that the integrator operates as a low-pass filter for input signals (Vi) and
works as a high-pass filter for the quantisation noise (N) by summing the error voltage.
The quantisation noise is spread to a higher frequency range, and it can be removed by a
digital filter afterwards, which achieves low noise at the final output. The purpose of the
feedback loop is to maintain the average output of the integrator near the reference level
of the comparator, which averages the DC voltage rather than filtering it out. Thus, the
delta-sigma modulator typically induces a chopper stabilisation module to cancel out the
DC offset variation [79].
+_
+
Vi(s) Vo(s)
N(S) quantisation noise
1
s
Figure 2.10: One-bit delta-sigma modulator in s-domain.
2.3.3 Neural Signal Processing: Spike Detection vs Spike Sorting
As the amount of brain data captured using microelectrodes increases, the transmission of
the recorded information poses a growing challenge for microelectronics. The bandwidth
of communication for implantable devices is fundamentally governed by the energy budget
and total amount of heat dissipation permitted to prevent tissue injury. These constraints
restrict the wireless transmission of such a large quantity of data. Consequently, it is
crucial to identify feasible routes for performing a portion of signal processing on-chip
while adhering to a minimal power budget. In this section, we review two processing
schemes, including spike detection only and spike sorting.
18 Chapter 2 Neural Information Processing
Spike Sorting: The raw data collected from neurons can be effectively filtered into local
field potentials (which reflect the dynamics of neural tissues around the electrodes and are
generated by the input currents of the dendrites of the surrounding neurons) and ’spike
trains’ [80]. The effect of filtering is to make spikes in background noise visible. The
filtered data is then separated into ’spikes’ and ’background noise’ (referred to as spike
detection) using methods such as establishing user-defined thresholds. The spike shape
is affected by a number of factors, such as the distance and orientation of the recording
electrode from the neuron [81]. However, to classify the shapes, several features, such as
spike amplitude and spike width, are selected, and based on these features, the spikes are
classified into ’clusters’ [82]. Each cluster is then linked to a single neuron, while clus-
ters that cannot be recognised are linked to ’multi-unit’ activity (typically characterised
by a lower amplitude) [66]. In summary, the flowchart for spike sorting consists of an
alignment of detected spikes, feature extraction, clustering, and classification, which is
the identification of clusters with original neurons.
Spike Detection: Spike detection is an essential technique that distinguishes spikes (real
brain activity/events) from background noise. Setting a threshold is the most typical
method for performing this operation. The threshold setting can be fixed or adaptable.
Fixed threshold utilises a constant circumstance throughout the duration of the recording.
The adaptive threshold strategy, on the other hand, monitors background changes and
modifies the threshold value accordingly. In any case, the threshold value must be selected
with care because, if the threshold value is too low, noise fluctuations may produce ’false
positive’ events, and if the threshold value is too high, spikes with low amplitudes will be
missed [83]. In the early days [84] of spike detection, the thresholds were manually set by
the user. Whenever the voltage signal exceeded the threshold, a pulse would be created
to represent an increase. If a spike waveform was required, specific user-specified points
would be collected after each threshold.
Selection: The decision between spike sorting and spike detection can be influenced by
the following factors:
Application: not all applications require precise spike sorting to reflect neuronal
activities. For instance, Kloosterman et al. [85] state that the signal ’read’ by cuff
electrodes derives from the multi-unit activity and that identification of individual
spikes is not necessary. A measure of overall activity over time frames longer than
individual spikes suffices.
Design constraints and trade-offs: It is of great interest to the neuroscience com-
munity to implement spike sorting algorithms on a low-power device in order to
enable wireless transmission of data and to provide ecologically and safely secure
settings for animal research. Wendler et al. [86] highlight the trade-offs that must
be addressed when creating a neural interface for the future generation. With the
consideration of the power budget of the system, it will be necessary to weigh the
Chapter 2 Neural Information Processing 19
application and scheme trade-offs carefully.
2.3.4 State-Of-the-Art Neural Front-Ends
The comparison among the state-of-the-art neuron front ends (from 2018 to 2022) and
the specification of our work is presented in Table 2.4. The majority of offset rejection
has shifted from AC-coupled to DC-coupled solutions in order to increase the recording
channel within a limited power and area budget. In the following sections, we reviewed
the AC-coupled front-end from [69] and a DC-coupled solution from [68] as examples.
AC-coupled neural SoC: it includes the functions of recording and detecting spikes
[69]. The analogue pixel comprises AC-coupled capacitive-feedback amplifiers, including
low-noise amplifiers (LNA) and variable gain amplifiers (VGA) and 10-bit SAR ADCs for
neural recording. Besides, the local digital processor conducts spike detection on-chip.
The architecture is presented in Figure 2.11 and the detailed performance is listed in
Table 2.4. The input AC-coupling capacitors are utilised to prevent the DC offset and
provide the proper DC level to the input transistor of LNA in order to bias it in the
correct operation region. In the feedback circuit, the pseudo-resistor determines the high-
pass filter cutoff frequency and the DC input bias of the pixel. The tunable pseudo-resistor
allows the cutoff frequency of the HPF for LFP (5 Hz) and AP to be adjusted (300 Hz),
which features the system with higher flexibility. However, compared with other works
with delta-sigma modulation [68;87], this offset rejection method still occupies a larger
channel area (shown in Table 2.4).
LNA VGA SAR
ADC
Spike
detection
Vin
VREF
Cin
Cf1
PR1
Cvar
Cf2
PR2
Figure 2.11: AC-coupled capacitive-feedback front-end.
Delta-sigma ADC: This front-end is fabricated under the electrode within an area of
70 ×70µm2is proposed in [68]. The presented first-order incremental ∆Σ ADC utilises
Gm-C integrator and current feedback. The architecture and its performance are shown
in Figure 2.12 and Table 2.4 respectively. The weak signal without pre-amplification does
not require strict linearity, and thus, it allows us to utilise the Gm-C technique at the
input, which can achieve higher area efficiency. As for the feedback path, the modulator
chooses the current domain, which can avoid adding extra input transistor pair into the
20 Chapter 2 Neural Information Processing
system. In order to reduce the area, this front-end enables three scale modes of ADC that
can cover the full-scale of ±11.25mV, ±22.5mV and ±45mV. The simple 1-bit ∆Σ ADC
with Gm-C input utilises the three different current sources to accommodate a wider range
of DC input voltage and occupies less area. In order to compensate for the DC offset,
it utilises a dynamic reference voltage VREF that is controlled by the digital module and
can cover the offset of up to ±200mV . It inspires us to design the ∆Σ ADC in our neural
front-end to achieve low noise, low area and low power consumption neural front-end in
Chapter 8.
Gm
1-bit
ADC
C
Vin
VREF
VBIASreset clk
VDD
I11 I22 I45
IFB Feedback (FB)
full scale
mode select
Vout
Figure 2.12: The architecture of one-bit delta-sigma ADC where utilises three scales
of current sources to conduct calibration.
2.4 Main Reference Circuit: Neural Pre-amplifier
Our group has proposed a neural front-end for spike detection which implements adaptive
thresholding by using memristive devices [30]. This design is referred to as the initial
front-end of this thesis. This section covers the design concept and basic operations of the
initial front-end. The author’s work on furthermore performance analysis is in Chapter 4,
and the following designs are in Chapter 5,6and 8.
2.4.1 Design Concept
It is important for implantable devices to increase the output data transfer rate and
decrease power consumption. Thus, we keep the necessary functions in implantable devices
and exclude other functions off-chip. It has been proved that it is sufficient to detect
neural spikes and transfer to off-chip [45] since most of the information representing the
extracellular activities is in APs. Therefore, we can include the main processing function
of spike detection and digitisation in the system. The duty cycle of APs is around 2% to
20%, which implies large potential power reduction when the system only processes and
Chapter 2 Neural Information Processing 21
Table 2.4: Comparison among state-of-art neural front-end and the specification of this work.
Parameter Dorigo 2018
[68]
Nikas 2019
[78]
Lee 2020 [88] Wang 2021
[79]
Yoon 2021
[69]
Wendler 2022
[87]
Target
Technology 180nm CMOS 180nm CMOS 110nm CMOS 55nm CMOS 65nm CMOS 180nm CMOS 180nm CMOS
+ Memristor
Functionality recording only recording and
detection
recording only recording and
detection
Coupling type DC DC DC DC AC DC DC
Architecture CT incremen-
tal ∆Σ
IA+CTDSM IA+CTDSM ∆Σ neural ampli-
fier + SAR
ADC
CT incremen-
tal ∆Σ
CTDSM
Area per channel
(mm2)
0.0049* 0.088 0.078 0.0077 0.0062* 0.0046* <0.005
Power (µW ) 46.29 23 6.5 61.2 2.72 (record-
ing module)
8.57 0.75
DC rejection (mV ) / 12.5 ±50 ±70 / 120 ±50
Bandwidth (Hz) 10k 5k 10k 10k 10k 10k 10k
FSR (mV )±45 26 300 148 / ±7.6 50
IRN (µVrms) 20.19 7.3 9.5 5.53±0.36 8.98 4.46 <10
*the area per channel of recording module only
22 Chapter 2 Neural Information Processing
transmits neural spikes and outputs redundant signal as zero/null [89]. Therefore, the
objective of the processing system is to detect neural spikes and output them as digital
signals.
However, the processing system is fed by the extracellular raw signal collected directly
from the neural interface, which mixes both local field potential and action potential. The
clean and stable AP contributes to higher accuracy of threshold detection of neural spikes.
Since the LFP fluctuates on a large scale and may submerge the desired AP in voltage
amplitude, we need to separate the AP from the raw signal from the aspect of frequency
as the first step.
With action potentials fed into the system, it conducts detection after amplification gen-
erally so that the signal occupies a large voltage range and is less sensitive to noise. The
aim is to output sparse neural spikes; thus, the amplification of discarded data points still
occupies considerable energy. Another challenge is that APs are in the minute amplitude
range that requires an expensive amplifier operating in low distortion and high sensi-
tivity. With the consideration that action potential is in low frequency and sustains in
milli-second, we can integrate the micron-volt signals to milli-volt slowly instead of purely
amplifying. As for spike detection, a simple solution is to conduct threshold detection
that biopotentials amplitude crosses the specified threshold voltage will be detected as a
spike, which requires a reference voltage input. Therefore, we proposed a fully differential
amplifier with appropriately sized load capacitors for enabling signal integration, which
is an input of a train of APs and a reference voltage at each amplifier input terminal,
respectively. Then, the output of the amplifier connects to ADC to realise digitisation,
which output can be transmitted off-chip for further processing and/or storage.
Besides, there are other tricky design concepts that access less power consumption and
high flexibility. Since input signals of interest are sparse, the system is not necessary
to operate on the whole process which can be controlled by fixed external signals or
asynchronous signals. In addition, the spikes provoked by the stimuli last for 1ms typically
[90]. Thus, this architecture is controlled by clocking signals, switching between ‘operate’
and ‘off modes, which is a solution to energy reduction. With regard to offset, we induced
memristive devices (whose resistive state can be programmed) in two current branches to
tune the offset instead of designing an expensive offset compensation circuit, making the
circuit more flexible.
2.4.2 Operational Principles
The schematic diagram is presented in Figure 2.13, which is a fully differential amplifier
wrapped by the dynamic latch comparator, while the biasing circuit unit can be shared
with multiple channels. The whole system mainly conducts the threshold detection for
APs and outputs digital signals where ones represent spikes and zero as non-spikes. The
Chapter 2 Neural Information Processing 23
following part will access to four operation phases in detail and explain the main compo-
nents of the circuit.
M7 M8
M0
M1 M2
M3 M4
M6M5
R1 R2R3
C1 C2
Vdd
M9
clk_ana
clk_anabar
clk_rst
ina inb
mida midb
drain_a drain_b
M10 M11 M12 M13
M14
M15
M16
M17
M18
M19
M24
M25
M20
M22
M21
M23
Vdd
Vdd Vdd
clk
clk
clk
clk
clk
outb outa
Figure 2.13: The architecture of pre-amplifier. It consists of three parts: (i) the core
integrating amplifier, (ii) the dynamic latch comparator and (iii) the current bias control
unit. Control signals, including clk,clk ana,clk anabar,clk rst are all assumed to be
generated by voltage sources which are strictly periodic.
A completed detection cycle consists of four phases, which have been shown in Figure 2.14:
(i) reset, (ii)integration, (iii) digitisation and (iv) off phase. The detailed operation of each
phase will be explained as follows:
In the reset phase (i) the core amplifier is on (clk ana,clk rst: high, clk,clk anabar:
low) and the load capacitors are discharged (Vmida/b = 0), so that voltage/current in core
amplifier is initialised and cleared before integration commences in the next phase.
In the integrating phase (ii) (clk ana: high, clk anabar,clk rst,clk: low) the reset tran-
sistors (M5&M6) are switched off and the currents flowing through the branches of the
core amplifier drain into the load capacitors. From a ‘large signal’ perspective, Vmida and
24 Chapter 2 Neural Information Processing
0 50 100 150 200 250 300 350 400
0
1
2
Voltage (V)
0 50 100 150 200 250 300 350 400
0
1
2
Voltage (V)
0 50 100 150 200 250 300 350 400
Time (ns)
0
1
2
Voltage (V)
(a)
V(drain_b)
V(midb)
(b)
V(outa)
V(outb)
(i)
(ii)
(iii)
(iv)
V(clk_ana) V(clk_anabar) V(clk_rst) V(clk)
(c)
Figure 2.14: Transient simulation of one detection cycle of the neural spike. One
detection with four phases completes in 350ns as an example, which has been labelled
and highlighted with different backgrounds. It contains (i) reset, (ii)integration, (iii)
digitisation and (iv) off phases. (a) presents the control of clocks; (b) compares the
amplifier output (midb) with the drain of the input transistor (drain b); (c) presents the
digital output from DLC. Input signals are set as: ina = 1V+ 50µV and inb = 1V.
Vmidb continuously increase during integration. In terms of ‘small signal’, Vmidamidb
increases with time and normal operation is maintained so long as the cascode transistors
M3&M4 remain in the saturation region. The voltage difference between nodes mida
and midb is impacted by the charging speed/current and integration time. Memristors
R1&R2 work as trimming devices and tune the offset of the core with very high sensitivity
(1µV/k shown in [30]). At the end of this phase, Vmida/b should be high enough to suc-
cessfully trigger the DLC and Vmidamidb should be as large as possible for maximising
gain.
In the digitisation phase (iii) (clk ana,clk: high, clk anabar,clk rst: low) clk goes high,
triggering the DLC to perform the conversion of Vmida/b into the final digital outputs. By
convention, we take the output from the branch where output ‘1’ represents a spike while
‘0’ represents the absence of a spike. Shortly after the decision is committed by the DLC,
the core amplifier is turned off as the system re-enters the off phase.
Finally, in the off phase (iv) (clk: high, clk ana,clk anabar,clk rst: low), the tail
current is cut off by setting clk ana to zero. The pre-amplifier is turned off and stops
recording neural signals. clk anabar is also deactivated (goes to high), thus preventing
the accumulated charge across the large gate capacitances of M1&M3 from draining away.
After access to the basic operation of one detection cycle, the test with the differential
input range of ±500µV is presented in Figure 2.15. Input A (ina) is slowly swept between
Chapter 2 Neural Information Processing 25
0 0.5 1 1.5 2
1.0999
1.09995
1.1
1.10005
1.1001
0
0.5
1
1.5
2
Voltage (V)
0 0.5 1 1.5 2
0 0.5 1 1.5 2
1.0999
1.09995
1.1
1.10005
1.1001
0
0.5
1
1.5
2
Voltage (V)
0 0.5 1 1.5 2
Time (ms)
+500μV
V(ina) V(inb)
V(outa) V(outb)
(b)
(a)
-500μV
1.1V
Figure 2.15: Pre-amplifier basic functionality test: Input A (ina) is slowly swept be-
tween [1V100µV , 1V+ 100µV ] over 2ms while the pre-amplifier is carrying out a
conversion every 10µs to detect the relationship between inputs A and B. Input B (inb)
remains stable at 1 Vthroughout. In this test, the amplifier was balanced (R1 = R2).
When Vina < Vinb, the left branch current is larger than the right branch current, in-
ducing Vmida Vmidb >0. The DLC captures this relation and generates binary signals:
Vouta = 1 and Voutb = 0, which appears in the bottom panel as a predominantly orange
output trace. Conversely, when Vina > Vinb,Vouta = 0 and Voutb = 1, which appears
as a combined orange/blue output trace. Note: this type of simulation can also be used
to test the offset tuning range and tuning sensitivity on the resistive state of memristive
devices. When R1> R2, Vina must be lower than Vinb to ensure a balanced output,
creating an offset, this is read in the output trace as an encroachment of the blue region
into the orange (and vice versa for R1< R2).
[1V500µV , 1V+ 500µV ] over 2ms while the pre-amplifier is carrying out a conversion
every 10µs to detect the relationship between inputs A and B. Input B (inb) remains stable
at 1 Vthroughout. In this test, the amplifier was balanced (R1 = R2). When Vina < Vinb,
the left branch current is larger than the right branch current, inducing Vmida Vmidb >0.
The DLC captures this relation and generates binary signals: Vouta = 1 and Voutb = 0,
which appears in the bottom trace as a predominantly orange output trace. Conversely,
when Vina > Vinb,Vouta = 0 and Voutb = 1, which appears as a combined orange/blue
output trace. This type of simulation can also be used to test the offset tuning range and
tuning sensitivity on the resistive state of memristive devices. When R1> R2, Vina must
be lower than Vinb to ensure a balanced output, creating an offset. This is read in the
output trace as an encroachment of the blue region into the orange (and vice versa for
R1< R2).
26 Chapter 2 Neural Information Processing
2.5 Summary
The development of the electrode techniques enables researchers to acquire a larger
amount of neural data. Therefore, the primary responsibility of the front-end designer is
to process the weak signals from the electrodes while adhering to low-noise requirements
and power and area constraints. It necessitates the selection of appropriate DC offset
rejection and signal processing techniques. In this instance, we use a memristive device
in the CMOS technique to compensate for the DC offset while simultaneously integrating
spike detection on-chip to realise a multi-functional front-end.
Chapter 3
Integrating Memristive
Technology into CMOS Design
Memristor is an emerging two-terminal device whose resistance state (RS) can be changed
by the applied bias voltage. It is a programmable and non-volatile device at the nanometer
scale that is compatible with CMOS [31;32]. Considering that the memristance is tuneable
continuously [27], it can be regarded as an analogue device that processes the continuous
signal. With these characteristics, it can be utilised as a trimming device with the feature
of programmable resistance [30]. In this project, CMOS integrated circuits with memristor
implants operate more flexibly because the RS can be altered by applying a bias voltage,
operating in a ‘write once-read many’ regime.
Previously, we only had the Python model for this device in the range of [20kΩ, 120kΩ],
which is used on our instrument measurement interface. The first step in incorporating the
memristive model into circuit design is to make the memristive Verilog-A model accessible
to EDA tools. It can be utilised to represent the behaviour of memristive devices through
physical equations, laying the foundations to enable the inclusion of these devices into
integrated circuits. This chapter focuses on the specific Verilog-A memristor model (in
the range of 20k to 120kΩ) that uses quadratic fitting as proposed in [36]. In this
project, I converted the mathematical model into a Verilog-A model and made it available
to Cadence, whose methodology refers to [37]. In this chapter, we cover the procedure
and significant parameters of the Verilog-A model. Additionally, the integration of the
memristor model into the Cadence Virtuoso design environment is documented, along
with the methodology for integrating memristors with CMOS designs.
27
28 Chapter 3 Integrating Memristive Technology into CMOS Design
3.1 Verilog-A Memristor Model
The Verilog-A model provided in [37] represents the behaviour of memristive devices
through physical equations. There are two mathematical models for our memristive de-
vice: exponential [37] and quadratic [36]. But we only had the exponential Verilog-A
model, which can reflect the physical device with higher accuracy. However, the RS is
limited to a narrow range of [10k,17kΩ]. The author applied the quadratic model to
CMOS design to access the wider RS range of 20k to 120kΩ.
1. module analytical (p, n);
2. inout p, n;
3. electrical p, n;
4. parameter real Ap = 0.12340;
5. parameter real tp = 2.74111;
6. parameter real An = -0.33000;
7. parameter real tn = 2.59685;
8. parameter real rp0 = -40928.13784;
9. parameter real rp1 = 55117.97865;
10. parameter real rn0 = 41366.35820;
11. parameter real rn1 = 7789.66771;
12. parameter real Rinit = 40000;
13. parameter real eta = 1;
14. parameter real ap=0.225;
15. parameter real bp=4.12;
16. parameter real an=0.2801;
17. parameter real bn=4.10;
18. real Rmp, Rmn, svp, svn, vin, RS, IVp, IVn, IV;
19. real first_it eration, R0_last, dt, it;
20. analog function integer stp;
21. real arg; input arg;
22. stp = (arg >= 0 ? 1 : 0 );
23. endfunction
24. analog begin
25. if (first_iteration==0) begin
26. it=0; R0_last=Rinit;
27. end
28. dt=$abstim e-it;
29. vin=V(p,n);
30. Rmp=rp0+rp1*vin; Rm n=rn0+rn1*vin;
31. if (vin>0) begin
32. svp=Ap*(-1+exp(abs(vi n)/tp));
33. RS=(R0_last+svp*Rmp*(Rmp-
R0_last)*dt)/ (1+svp*(Rmp-R0_last)*dt);
34. end
35. else begin
36. svn=An*(-1+exp(abs(vin)/tn));
37. RS=(R0_last+svn*Rmn*(Rmn-
R0_last)*dt)/ (1+svn*(Rmn-R0_last)*dt);
38. end
39. if (RS>=Rmp && vin>0) RS=R0_last;
40. if (RS<=Rmn && vin<0) RS=R0_last;
41. if (abs(vin)<=0.5) RS=R0_last;
42. IVp=ap*(1/RS)*sinh(bp*vin);
43. IVn=an*(1/RS)*sinh(bn*vin);
44. IV=IVp*stp(vin)+IVn*stp(-vin);
45. I(p, n)<+ IV;
46. R0_last=RS;
47. first_iteration=1;
48. it=$abstime;
49. end
50. endmodule
Listing 1: the Verilog-A memristor model representing the in-
house fabricated Pt/TiOx/Pt device using quadratic fitting
In this chapter, we convert the quadratic
mathematical model in [36] into the
Verilog-A model (in Listing 1), whose
methodology refers to [37]. The explana-
tion and implementation of the Verilog-A
model in Cadence Virtuoso are given as fol-
lows.
As a two-terminal device, p, n is de-
fined as the ‘inout’ ports of the mem-
ristor, where the applied bias voltage
(vin) can be used for calculation and
the current through the device will
be output (lines 1-3, 29, 45). The
RS can be accessed through the non-
linear IVcharacteristic in lines 42-
44.
The fitting parameters, switching pa-
rameter (eta) and initial RS (Rinit)
are defined from lines 4 to 17. The
fitting parameters are extracted from
the experimental results of an in-
house fabricated device with a spe-
cific stimulus, while the details can
be found in [37]. The switching di-
rection parameter (eta) is 1 in this
model, indicating the direction that
the positive voltage induces higher
RS. User-defined initial values for
RS should be within the proper
range determined in line 30.
Lines 20-23 define a ‘step function’, which is then employed in line 44 to determine
Chapter 3 Integrating Memristive Technology into CMOS Design 29
the output current branch. For instance, when vin is positive, the result of ‘stp(vin)’
is one while ‘stp(-vin)’ is zero; thus, the current branch ‘IVp’ will be output.
Analog block (lines 24-49) indicates two features of the device operation: iteration
and transient. Lines 25 to 27 present the setup of the first iteration. In this case, the
start time is set to zero and the latest RS is taken from the defined parameter Rinit.
The time step can be obtained from lines 28 and 48 by subtracting the absolute time
from the reference time. After deriving RS (lines 31-41), the current through the
device will be updated (lines 42-44) and passed to ports (line 45). Then, the latest
RS and absolute time will be updated for the next iteration (lines 46-48).
The boundaries of RS can be derived from line 30. Within the RS boundaries,
RS can be calculated under the condition of constant bias voltage in lines 31-38.
However, the switching fails to realise if the latest RS exceeds the boundaries or the
bias voltage drops below 0.5V(lines 39-41).
In conclusion, Verilog-A models can only be used to simulate transients and can be used
to 1) conduct static current-voltage measurements within specific boundaries; 2) gather
transient switching characterisation by applying bias voltages to initial RS. In order
to process the above utilisation, we need to set up two stages: 1) read RS by applying
triangular pulse with 0.5Vamplitude, which prevents the device from switching; 2) write,
or we said, change the RS, by applying pulses with defined duration/width, polarity,
amplitude, and numbers of pulses. For extending the application, fitting parameters can
be changed to represent other memristive devices. At this stage, the model does not
incorporate AC analysis/small-signal modelling, noise performance, parasitics and device
variation.
3.2 Integration of Memristor Model into Cadence
3.2.1 Import Verilog-A Model into Cadence
In order to utilise the device in circuit design as well as simulation, cooperating with
electronic devices, we develop a symbol that links to the Verilog-A model so that it can
be incorporated into circuits as an element. The instruction is shown as follows:
Build the library and refer to chosen technologies. In order to cooperate memristor
with CMOS, users are supposed to be familiar with the behaviour of the memristor
quantitatively, such as the range of high/low resistive state, the allowed range of
applied voltage/current, static IVcharacteristics and etc., which allows them to
choose a more suitable technology, as well as define circuit specification with more
realisable. In this case, we integrate the Verilog-A model in Cadence and choose
30 Chapter 3 Integrating Memristive Technology into CMOS Design
0.18um technology with four fabrication metal masks for demonstration, where the
proposed memristor was fabricated with Metal4 as ports.
In library Manager, we create a cell view in a specific library with Verilog-A type,
named ‘memristor’ in library ‘DesignMethodology’ (Figure 3.1(a)). Then, we paste
the provided Verilog-A code into the text editor and save the file.
From the toolbar in the text editor, we can access create ‘cellview from cellview’ to
create a symbol from Verilog-A memristor model (Figure 3.1(c).
We assigned the position of ports on the left and right sides which can be auto-
matically detected from the Verilog-A code, followed by the symbol design (Fig-
ure 3.1(b)).
The memristor model can be applied to the schematic, operating with transistors,
resistors, capacitors and etc. Besides, the model can be utilised flexibly, where fitting
parameters (shown in Figure 3.1(d)) can be modified to represent other memristor
models. In this case, we only have access to construct a single proposed memristor
with voltage sources to investigate the switching characterisation.
(a) (b)
(c) (d)
Figure 3.1: The operation sequence of importing memristor into Cadence. (a) Process
of building a new cell view in Verilog-A type for memristor. (b) Symbol of the memris-
tor. (c) Process of creating a symbol for memristor from Verilog-A cell view. (d) The
parameters of the memristor can be modified in the object properties window.
3.2.2 Simulation Setup for Memristor-based Circuit
Considering that the Verilog-A memristor model calculates RS against time and keeps
tracking the change of RS, the model is supposed to run in transient simulation to process
Chapter 3 Integrating Memristive Technology into CMOS Design 31
both write and read operations. In this case, we took a single memristor with voltage
sources as a simulation example that was provoked by a chain of pulses followed by
a triangular wave to conduct the write and read of the memristor, respectively. The
operation steps will be shown as follows:
(a)
Time
Voltage
t
w,in
t
w, R
V
b
V
read
V
read
(b)
Figure 3.2: (a) Testbench of applying pulse and triangular wave to memristor to con-
duct the write and read operations in the transient simulation. (b) Pulse chain that
provokes the memristor. The input signal sequence can be divided into two parts: trian-
gular wave and pulse. For all the simulations of our device, the read voltage is defined
as Vread = 0.5Vwith 1ms duration. The pulses can be determined with specific dura-
tion/width (tw,R), amplitude (Vb) and the number of pulses to provoke the memristor.
With the combination of two stages, the RS of the memristor model can be tracked for
each stimulus.
In this demonstration, we built the schematic with memristor, piece-wise linear
(PWL) and pulse voltage sources (Figure 3.2(a)). In the schematic, we have de-
fined the direction that voltage sources are connected to ‘positive’ (p) port of the
memristor. The positive bias voltage (Vb>0) from ‘p’ provokes the memristor to
higher RS, while the RS will decrease when positive voltage applies at ‘n’ port. In
testbench (Figure 3.2(a)), the proposed model is in OFF transitions under positive
voltage, while negative voltage leads to ON transitions. Two voltage sources gen-
erate triangular waves and pulse alternately to process that the pulse changes RS
of memristor followed by a triangular wave which keeps tracking RS. The detail
of the input signal has been shown in Figure 3.2(b). In this measurement for the
proposed model, the identical triangular wave is defined as 0.5Vpeak amplitude
(Vread) within 1ms duration. The pulse mainly has three variables: width (tw,R),
amplitude (Vb) and the number of pulses.
After setting up the testbench and running the transient simulation, we are supposed
to record the change of RS. In this case, we need to record every current across
the memristor at Vread = 0.5V. With both read voltage and current, the RS can be
obtained by dividing Vread = 0.5Vby read current. For detail operation, users can
1) process the division in Cadence calculator and plot RS; 2) send the RS to the
table and save it as a ‘.csv’ file; 3) process the data in Matlab in order to extract
and plot the RS at Vread = 0.5V.
32 Chapter 3 Integrating Memristive Technology into CMOS Design
The simulation example of the Verilog-A model in Cadence is given in Figure 3.3. This
type of simulation allows us to elaborate on the effect of the resistance state from the
applied pulse width, amplitude and polarity.
1s
ss
(a)
(b)
Figure 3.3: Simulation results for the quadratic memristive Verilog-A model in Cadence.
(a) shows the responses of the model to 1,10,100µs pulse widths with constant amplitudes
(±2V). (b) presents the model response to the voltage ramp from 1V to 3V.
3.3 Summary
This chapter shows the behaviours of memristor from explaining the Verilog-A model in
terms of non-linear IVrelationship and RS switching due to bias voltage. Besides,
the methodology of utilising the memristive device into Cadence for circuit design and
simulation is given. It can be regarded as the foundation for the below chapters that
apply the memristor to the amplifier, including defining and measuring the performance
of the proposed amplifier.
Chapter 4
A Memristor-based Pre-amplifier
Topology
Based on the design requirements and application provided in Chapter 2, this chapter
analyses the pre-amplifier at the transistor level and provides theoretical calculation of
the core memristive OTA. Since the proposed OTA operates in a ’start-stop’ scheme, its
performance metrics are slightly different from the conventional OTA. And the differences
and explanations are given below. In this thesis, we modified the initial front-end (Fig-
ure 2.13) and the proposed schematic is shown in Figure 4.1. In the initial front-end [30],
the clocking signal (clk) is generated by the output of core OTA (mida). On the other
hand, the clk in the proposed OTA is set to be generated by the external clocking module,
which guarantees the DLC capture the sufficient voltage difference from the core OTA.
The next chapter presents a transient simulation for signal detection with more details.
4.1 Transistor Level Analysis
After introducing basic functions of the pre-amplifier in Chapter 2, we analyse the be-
haviour of significant transistors of memristive OTA for afterwards design. The dynamic
OTA operates in four phases: i) initial, ii) integration, iii) digitisation and iv) off.
The input differential pair (M4&M5) provides the main transconductance for the core am-
plifier which converts the effect of gate voltage into current in CMOS device. To achieve
maximum gain, the gmneeds to achieve the highest value (for a given bias current), which
indicates the input transistors need to bias to the subthreshold region. In addition, a max-
imum gmmakes a contribution to reducing thermal noise component of the input-referred
noise and the use of large input transistors achieves a lower flicker noise component.
The cascode transistors (M6&M7) in the core amplifier mainly contribute to increasing
output resistance and being shield devices to isolate the input pair from memristive devices
33
34 Chapter 4 A Memristor-based Pre-amplifier Topology
I
II
III
M1 M2
M3
M4 M5
M6 M7
M9M8
R1 R2
R3
C1 C2
Vdd
M10
clk_ana
clk_anabar
clk_rst
M11 M12 M13 M14
M17
Vdd
outaoutb
clk clk
clk
ina inb
mida midb M18
M16
M15
M19
drain_a drain_b
Figure 4.1: The schematic of the proposed pre-amplifier. Compared to the initial
design in Figure 2.13, this circuit uses external clk to control that the DLC can be fed
by sufficient voltage difference and increase detection accuracy.
and dynamic voltage changing due to integration. Thus, drains of input pair (drain a/b)
can be maintained at a constant voltage level, which avoids fluctuating input pair.
The memristor, along the current branch, plays as a trimming device in the core amplifier.
Generally, the circuit requires an adjustment circuit to cancel or compensate the inherent
offset voltage to prevent drift after fabrication. But in this case, we do not intend to
cancel the inherent offset; on the other hand, we take advantage of the inherent offset
as part of reference/threshold voltage during threshold detection. The differential output
Vmid =Vmidb Vmida is impacted by the inclined current with the same integration time.
Chapter 4 A Memristor-based Pre-amplifier Topology 35
We applied memristors (R1&R2) that can be programmed to trim the effective resistance
which inclines the balance of two current branches. Thus, it requires that the tuneable
range should cover the range of Vspike,max ±|Vos,max|, where Vspike,max is the maximum
potential of neural spike and Vos,max is the maximum offset voltage of the core amplifier.
As for the bottom transistors (M8&M9), they operate as switches rather than active
loads. During the ‘initial phase,’ the bottom transistors are activated to discharge the
load capacitors (C1&C2) and initialise the status of the input and cascaded pairs. And
they are turned off during the ‘integration phase’ to guarantee that charges flow into the
capacitors to realise integration and amplification. As a result, the design of the bottom
transistors should prioritise minimising parasitics and leakage current.
In this section, we highlight some significant components and preliminary design consid-
erations. The detailed theoretical analysis and calculation of specified indicators are given
below.
4.2 Theoretical Analysis of Performance Metric
Considering that the core amplifier is a dynamic amplifier where the integrated output
signals (mida/b) keep changing the DC operation point, the performance of the amplifier
needs to be re-defined based on the transient analysis since the traditional DC/AC analysis
based on the static operation point is unsuitable for this case. In this section, we assess
the foundation of the performance metric and tailor it to the integrating amplifier. Most
of the indicators mainly depend on the ‘integrating phase’ (iii in Figure 2.14) when the
amplification is processed. Thus, the analysis will be focused on the core amplifier in the
integrating phase.
Furthermore, the bias current (Itail) distributes into differential branches with a slight
incline (∆i) depending on weak input signals. This allows us to conduct both ‘large’ and
‘small’ signal analysis by using transient simulations to obtain the relevant data. Thus, we
define the Vmida,Vmidb and Itail as ‘large’ signals in this case, while the voltage difference
Vmid =Vmidb Vmida and inclined current ibelong to ‘small’ signal. The following
access to each performance indicator one by one with the insight of definition foundation
and the derivation of tailored definition. This chapter mainly focuses on the performance
analysis of core OTA, whose parameters are listed in Table 4.1.
4.2.1 Overview of Performance Metric
The indicators are divided into different types in Table 4.1. Moreover, we briefly de-
scribe significant indicators and provide a detailed explanation for the specific ones in the
following sections.
36 Chapter 4 A Memristor-based Pre-amplifier Topology
Table 4.1: Specification of the memristive integrating OTA.
Type Parameter Target
DC
Bias current Input common mode range VIC R 50mV
Input differential voltage range VIDR 5mV
Offset compensation by memristors VCOMP VOS +500µV
AC
Open loop voltage gain AOL 26dB
Bandwidth BW 10kHz
Power supply rejection ratio PSRR 100dB
Common mode rejection ratio CMRR 100dB
Transient Sampling rate Fs20kHz
Noise Input-referred noise en10µVrms
Minimum ratings Power dissipation 750nW
DC There are two types of input ranges: i) common mode range and ii) differential
voltage range. The input common mode range VIC R is defined as the voltage that keeps
the input transistors in the subthreshold region, which analysis is in section 4.2.2. The
input differential voltage range VIDR is relative to the linearity where the output is linear
to the input voltage. Due to the electrode can induce up to 50mV DC offset to the
amplifier, to guarantee the input transistors are in subthreshold region, VICR is required
to be larger than 50mV. As for the differential input, it contains both LFP and AP that
we need to set VICR VLF P,max +Vspike,max 5mV .
The other highlighted parameters are relevant to offset. The inherent offset voltage is
dependent on the sizing and layout design. And we determine that the compensated
voltage by memristor VCOM P needs to cover both the spikes and inherent offset of the
circuit (Vspike,max ±|Vos,max|), where the amplitude of neural spike is Vspike,max = 500µV .
AC Due to the amplification being realised by integrating the capacitor, the open loop
voltage gain AOL/output voltage mid should be relative to the accumulated charges
in the capacitors. Furthermore, the detailed calculation is in section 4.2.3. We required
the OTA to amplify the minimum spike 50µV to millivolt-level, which requires 20V/V
(26dB) amplification. As for the frequency response, the bandwidth is set in the range
that covers the neural signals which are up to 10kHz. Due to the front-end being required
to reject the DC offset, the OTA should operate as the bandpass filter by applying it into
the open/closed loop network as an AC-coupled front-end. In addition, the OTA can be
applied to the DC-coupled solution described in Chapter 6.
Transient The maximum frequency of the neural spike is 10kHz. According to the
Nyquist theorem, the sampling frequency must be at least twice the signal of interest; we
set the sampling frequency is required to be higher than 20kHz.
Noise Noise from the input amplifier is typically lower than that from electrodes [91].
Moreover, the noise introduced by the electrodes is typically 10µVrms [13]; thus, we target
Chapter 4 A Memristor-based Pre-amplifier Topology 37
the input-referred noise ento be lower than 10µVrms.
Power consumption As estimated in Chapter 1, the maximum power consumption of
the implanted integrated circuit is restricted to 15mW. Moreover, we estimated that the
transmission module occupies 50% of the power budget; the 10,000-channel front-end can
only consume 750nW per channel.
4.2.2 Input Range
As the analysis above, the input transistors M4&M5 are biased in the subthreshold region
to achieve higher voltage gain and lower input-referred noise. It implies two operating
conditions: (1) A minimum drain-source voltage |Vds,min|= 3·VT, where VTis the thermal
voltage and good rule of thumb for ensuring in subthreshold region [92]. (2) the gate-
source voltage |Vgs,4|(<|Vth,4|) that allows the transistor to pass Itail/2in subthreshold
region. These two conditions can be unfolded with the aid of Figure 4.2 as follows.
M3
M4 M5
M6 M7
R1 R2
Vdd
clk_anabar
ina inb
mida midb
Vbias
drain_a
Figure 4.2: Part of schematic of core amplifier aiding the input range analysis.
Looking into Figure 4.2 from VDD to the input, the common mode input voltage VCM
reaches the top bounded:
VDD |Vds,sat,3|−|Vgs,4| VC M (4.1)
where Vds,sat,x is the drain-source saturation voltage of transistor x. Exceeding the bound-
ary causes M3 to triode and simultaneously encroach on Vgs,4, progressively shutting the
amplifier down.
The bottom boundary hinges on maintaining the input differential pair in the subthreshold
region (|Vds,4| |Vds,min,4|= 3 ·VT):
38 Chapter 4 A Memristor-based Pre-amplifier Topology
|Vds,4| (VCM +|Vgs,4|)(Vclk anabar +|Vgs,6|)3·VT(4.2)
where Vgs,6is the gate-source voltage allowing the cascode transistor to pass Itail/2.
This is also treated as approximately constant in this analysis. Under normal operation,
the second term is recognised as Vdrain a. This unfolds to:
VCM Vclk anabar +|Vgs,6|−|Vgs,4|+ 3 ·VT(4.3)
Here, the cascode transistor M6 enforces a specific and relatively fixed value of Vdrain a
under the control of Vclk anabar (similarly for M7 and Vdrain b). Combining Equations 4.1
and 4.3, we can find the approximate value of Vclk anabar above which the input differential
pair runs out of the common mode range:
Vanabar,low =VDD |Vds,sat,3| 3·VT |Vgs,6|(4.4)
We can see the trade-off between common mode and integration voltage ranges (directly
connected to gain), depending on the clk anabar. If Vclk anabar increases, push the bottom
boundary of VCM to narrow the common input range; meanwhile, it allows large headroom
for the signal integration at nodes miba/b which boosts the voltage gain.
4.2.3 Open Loop Voltage Gain
For a fully differential amplifier, the voltage gain is defined as the ratio of the output am-
plitude difference over the input amplitude difference: Vout /Vin. As for the proposed
amplifier, the voltage gain will be translated into Vmid/Vin, where Vmid is captured
at the end of integrating phase and Vin is the input difference between APs and refer-
ence voltage in practice. The APs are in low frequency that each spike features of the
order of 100smicron seconds, while a completed detection cycle lasts within hundreds of
nanosecond levels, which implies the input voltage can be regarded constant for analysis.
First, we go through the signal conversion and flow within the core amplifier during the
integrating phase step by step. In this structure, the input pair (M4&M5) contributes
the main transconductance to convert the voltage into current through the device. The
cascode pair (M6&M7) mainly contributes to output resistance and shields the voltage
fluctuation from nodes mida and midb, maintaining the stability of input transistors.
Then, the branch current charges the load capacitors that induce both output voltage
Vmid and the voltage difference Vmid increases gradually, where converts the current
back to voltage and realises amplification. The process of deriving the formula will be
given as follows.
Chapter 4 A Memristor-based Pre-amplifier Topology 39
The input signal and reference voltage are input into the amplifier through M4&M5
where the voltage difference Vin is converted into current difference i. The conversion
degree depends on the transconductance gmof the input pair, which represents the signal
amplification ability. The current difference will be:
i= Vin ·gm(4.5)
As the branch current drains into load capacitors within the same period, the charges in
the capacitor will be accumulated and the charge difference Qwill be increased during
this process. The difference of charge on the load capacitors:
Q= i·τ(4.6)
where τis the integration phase duration. Finally, this gets transformed into the voltage
difference we observe at Vmid through the load capacitance C:
Vmid = Q/C (4.7)
Combining the above yields the gain (G):
G=Vmid
Vin
=gm·τ
C(4.8)
From the view of the ‘large’ signal, currents filling each load capacitor are approximately
constant and equal, and the output Vmid is mainly dependent on the half-tail current
(Itail/2) integration within τ:
Vmid =Q
CItail/2·τ
C(4.9)
where Qis the total charge accumulated on each node (mida/b) as a result of the tail
current. Since the load capacitor has discharged clearly in the reset phase, we induce Qto
represent the accumulated charge during integration. Similarly, Vmid represents the ‘large’
voltage signal at node mida/b during integration. We can easily get that the integration
time τis mainly dependent on the ‘large’ signal integration, and τcan be expressed as:
τ=Vmid ·C/Itail/2(4.10)
Then, we can substitute Equation 4.10 into Equation 4.8 and obtain gain (G) as:
40 Chapter 4 A Memristor-based Pre-amplifier Topology
G=gm·Vmid
Itail/2
(4.11)
We can divide the formula into two parts: the transconductor efficiency factor of the
input transistor pair (gm/Itail/2) and the ‘large’ signal output voltage (Vmid), both of
which depend on the differential gain of integrating amplifier. The maximum gm/Itail/2
can be obtained when the input transistors are biased in the subthreshold region. As for
the output voltage Vmid, it needs to be integrated large sufficient to boost the gain, while
Vmid should be lower than the voltage that forces the cascode pair (M6&M7) into triode
region.
4.2.4 Noise Performance
The input-referred noise derivation in continuous mode is given in this section. The
standard MOSFET input-referred noise model containing both thermal and flicker noise
is given by the following expression for spectral density [93]:
V2
in std(f) = 4k T γ 1
gm
+K
CoxW Lf (4.12)
where kis Boltzmann’s constant, Tis the absolute temperature, γ=2
3for long-channel
transistors and higher for shorter channel devices, Ka typically empirically determined
factor scaling 1/f noise, Cox the gate capacitance, W, L the transistor sizes and fdenotes
(linear) frequency.
The analysis of each component in the core amplifier on noise contribution during the
integrating phase will be conducted. The criterion is whether the transistor contributes
to AC signal increment from input to output. M3 is beyond the differential branch and
mainly contributes to common-mode noise. The input pair (M4&M5) induces the primary
increment on the AC signal; thus, it is an effective noise source at mid nodes. The cascode
pair (M6&M7) operates as a ‘common-gate’ that contribute zero increments on AC signal,
which can be omitted in noise analysis. Finally, the reset transistors (M8&M9) are off
during the integrating phase. Similar to continuous mode amplifiers such as the Harrison
[94]: the input differential pair provides substantial gain through its gm, mitigating the
input-referred contributions from downstream elements (primarily the cascode transistors
and the memristive devices). Therefore, in a fully differential amplifier, the input-referred
noise will be doubled and given by:
V2
in(f) = 2 ·(4kT γ 1
gm
+K
CoxW Lf ) (4.13)
Chapter 4 A Memristor-based Pre-amplifier Topology 41
where all the component parameters in terms of gm,K/CoxW L represent the input tran-
sistors M4&M5.
4.2.5 Tuneable Range and Sensitivity
The memristive devices applied in the current branches regulate the charging speed to
load capacitors by modulating the effective output resistance of the core amplifier as seen
by the capacitive load. The calculation of branch resistance will be divided into two parts:
(1) find out the impedance of a drain-degenerated of M6, looking into the source of M6,
which schematic has been shown in Figure 4.3; (2) apply the drain degeneration analysis
again on the source of M4 (Figure 4.1) to calculate the branch impedance.
M6
clk_anabar
R1
Vtest
Itest
(a)
R1
Ro6gmVsg
Vtest
Vsg
+
_
Itest
(b)
Figure 4.3: The circuit diagram of M6 and R1 with a test voltage source presented as
(a) a large signal and (b) a small signal. Rorepresents the drain-source resistance.
To obtain the impedance, a test voltage source is applied on the source of M6 in Figure 4.3,
with the assumption of Itest. The effective impedance (Zs6) can be obtained by: Vtest /Itest.
Looking into the current path from the voltage source to R0, R1 and GND, we can obtain:
Vtest =Itest ·R1+ (Itest gm6Vsg,6)·Ro6(4.14)
To obtain the item Vtest/Itest , Equation 4.14 can be translated into:
Zs6=Vtest
Itest
=R1+Ro6
1 + gm6Ro6R1+Ro6
gm6Ro6
=1
gm61 + R1
Ro6(4.15)
where the approximation is induced due to in saturation operation: gm6Ro1. Zs6is
the impedance looking into the source of M6, gm6is the differential transconductance of
M6, and Ro6the output resistance of M6.
42 Chapter 4 A Memristor-based Pre-amplifier Topology
Extending this principle to calculate the impedance of M4, as drain-degenerated by the
M6-R1 cascade we obtain:
Zs41
gm41 + Zs6
Ro4(4.16)
which eventually unfolds to:
Zs41
gm4
+1
gm6gm4Ro4
+R1
gm6Ro6gm4Ro4
(4.17)
A similar expression also applies for the right current branch.
Setting A=1
gm4+1
gm6gm4Ro4and B=1
gm6R06gm4Ro4we can express the impedances seen
by M3 looking into each current branch as:
ZlA+BR1(4.18)
ZrA+BR2(4.19)
where Zl=Zs4is the left current branch impedance and Zris the right branch impedance.
Next, examining the distribution of tail current across the branches we obtain an expres-
sion for the left branch current ilas follows:
iliT
A+BR2
2A+B(R1+R2)(4.20)
where iT=i3is the tail current. Given that B1 (as it is the product of two maximum
FET amplifier gains), ilcan be further approximated as follows:
iliT
21B
2A(R1R2)(4.21)
Similarly for the right branch current ir:
iriT
21 + B
2A(R1R2)(4.22)
This yields a total current imbalance of:
iliri=iT·B
2A(R1R2) (4.23)
which if divided by the common transconductance of the input differential pair transistors
yields the required voltage offset to rebalance the branches as a function of the difference
Chapter 4 A Memristor-based Pre-amplifier Topology 43
in memristor resistive states:
Vos Vina Vinb =i
gm4,5
(4.24)
which when fully unfolded yields:
Vos (R1R2)iT
2Ro,casgm,in (1 + gm,casRo,in )(4.25)
where we have renamed our variables to explicitly stress the common values of output
impedances and differential transconductances of the input differential pair and cascode
transistors (Ro,cas = output impedance of cascode transistor, gm,in = transconductance
of the input differential pair).
Overall, Equation 4.25 shows that in small-signal conditions, the offset voltage of the
core amplifier is proportional to the difference in memristor resistive states divided by the
maximum transistor gains of the input diff pair and cascode transistors. This division
explains the extreme fineness of tuning achievable.
The tuning range can, in principle, be extended under the rule of Equation 4.25 for as long
as the underlying assumptions hold. We note two important limiting conditions: 1) If the
current imbalance becomes large, the assumption of equal gms on both current branches
collapses. Strictly when this occurs depends on the tightness of the specifications. 2) If
the voltage dropped across the larger of the pair R1,2becomes comparable to the capaci-
tor voltage range through which the amplifier can integrate while maintaining transistor
saturation (regular operation), eventually, the amplifier will run out of integration voltage
headroom.
4.3 Summary
This chapter accesses to transistor-level analysis for the proposed pre-amplifier and perfor-
mance metric theoretically with equations. Since the proposed memristive OTA operates
dynamically, some indicators differ from the conventional static OTA. We highlighted and
explained simulations to verify the differences and these in the next chapter.
Chapter 5
Methodology of Measuring
Memristor-based AC-Coupled
Pre-amplifier
After defining and analysing the circuit’s performance, the suitably defined performance
parameters from the previous chapter are assessed for an example design in simulation.
Since the capacitor integration and the control signal design are time-varying, the amplifier
evaluation is based on transient simulation and periodic steady-state (PSS) analysis rather
than direct DC/AC analysis. The overview of the simulation setup is listed in Table 5.1.
For each performance, we develop a simulation setup and analysis. Additionally, power
consumption is discussed separately. Pre-amplifier was applied to an open-loop network
(Figure 5.1) to reject the DC offset from electrodes. For these simulations, we used a
commercially available 0.18µm CMOS technology with VDD = 1.8V. The simulation of
this chapter is run based on the schematic in Figure 5.1, where the detailed OTA and
DLC designs are in Figure 4.1 and the sizes of transistors are in Table 7.1. A description
of the methodology and simulation setup will be followed by an evaluation and design
consideration based on the architecture.
5.1 Transient Simulation Setup
As mentioned above, the direct DC/AC simulation cannot be applied to this dynamic
amplifier. DC indicators, transient indicators and power consumption can be obtained
by utilising transient simulation in this design. Thus, the following provides a timing
diagram of the whole detection cycle at first to present the operation and function of
the circuit before analysing detailed performance. The timing diagram of one detection
cycle is presented in Figure 5.2, and the transient simulations are based on this operation
scheme.
44
Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 45
Ci
Ci
Vcm
Vcm
Rf
Rf
Gm
CL1 CL2
DLC
outa
outb
mida
midb
Vina
Vinb
clk_ana
clk_anabar
clk_rst
clk
Figure 5.1: Block diagram of the proposed neural recording front-end. The open-loop
network OTA filters the raw data to obtain the MUA. The minute signals charge the load
capacitors (CL1,2) to realise the integration and boost the voltage difference of mida &
midb, which triggers the DLC to process digitisation.
Table 5.1: Specification and overview of simulation methodology for the memristor-
based OTA.
Type Parameter Target Simulation
DC
VIC R 100mV
TransientVIDR 5mV
VCOMP VOS +500µV
AC
AOL 26dB
PSS
BW 100Hz-10kHz
PSRR 100dB
CMRR 100dB
Transient Fs20kHz Transient
Noise en10µVrms Transient/PSS
Minimum ratings Power dissipation 750nW Transient
5.1.1 Input Range
In order to experimentally demonstrate the input range of the amplifier, we performed
a series of experiments querying different potential range limitation factors in practice.
First, we checked the system’s behaviour at different stages as a function of common
mode voltage by running a series of integration cycles whilst sweeping VCM from 0V to
VDD in steps of 0.1V. At each run, the differential input was 50µV and the outputs were
registered after integrating for 70ns. Results were registered at: i) Vmidb, ii) Vmid and iii)
the overall system output after the DLC. Results are shown in Figure 5.3. Note: To check
for possible input signal history dependence during these tests, three integration cycles
with VCM = 1.8Vpreceded each test integration cycle. We have sample-tested a few runs
with initial VCM between 0.1V and 1.8V and confirm that the history-dependence effect
is negligible.
From the results in Figure 5.3 we can draw three key conclusions: 1) The DLC successfully
triggers for VCM between zero and 1.3V. This means that Vmidb is sufficiently high for the
46 Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier
0 50 100 150 200 250 300
0
1
2
Voltage (V)
0 50 100 150 200 250 300
0
1
2
Voltage (V)
0 50 100 150 200 250 300
Time (ns)
0
1
2
Voltage (mV)
0 50 100 150 200 250 300
Time (ns)
0
1
2
Voltage (V)
(i)
(iv)
(ii)
(iii)
(iv)
V(clk_ana)
V(clk_anabar)
V(clk_rst)
V(clk)
(a)
(b)
(c)
(d)
V(midb)
V(drain_b)
V(midab)
V(outb)
V(outa)
Figure 5.2: Transient simulation of one detection cycle of the neural spike. One de-
tection with four phases completes in 250ns, under the condition of Vina = 1V+ 50µV
and Vinb = 1V. It contains (i) reset, (ii) integration, (iii) detection and (iv) off phases.
Full detection cycle is presented, including (a) control signals, (b) output of the amplifier
(midb) and drain voltage of input transistor (drain b), (c) differential output of the OTA
(∆Vmid) and (d) digital outputs.
DLC to settle to an output within 30ns of its triggering (which occurs when clk goes high).
2) In this case, the DLC provides the correct answer so long as it triggers, but this might
change towards the edges of the range once we consider noise. 3) The actual analogue
gain of the amplifier remains close to the maximum (26dB) within a narrower region:
[0.7,1.2]V. We recommend that the maximum gain area be taken as the effective VCM
range to maximise the chances of correctly capturing small differential inputs under noisy
conditions. Nevertheless, this shows that by de-rating the specification of the amplifier to
higher Vmid we can extend its effective input range.
In order to visualise the effects leading to loss of gain outside the region VCM [0.7,1.2]V
we ran some unrestricted integration tests as shown in Figure 5.3 for different values of
VCM . The results are shown in Figure 5.4 where we observe that for VCM between 0.8V
and 1.2Vthe integration traces follow each other very closely, with traces at 0.7Vand
1.3Vbeginning to show more substantial deviations. We note how excessively low VCM s
Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 47
(c)
(b)
(a)
V(outa) V(outb)
Figure 5.3: Input range results of pre-amplifier. In this simulation, the common mode
voltage was swept from zero to 1.8Vwith 0.1Vdifferential input. For VCM [0 1.3]V,
we notice that (a) Vmidb reaches sufficiently high voltage to prompt a stable output from
the (c) DLC for our chosen differential input the output is always correct. However, the
core’s analogue gain in (b) is maximised in the narrower range [0.8,1.2]V.
shorten the peak without shifting (a result of desaturating the input differential pair but
not changing the integration range) whilst excessively high VC M s shift the peak without
changing its magnitude.
CM = 0.7, 0.8, 0.9, 1.0, 1.1, 1.2, 1.3V
Figure 5.4: Intermediate differential output Vmid evolution as a function of VC M .
Differential input voltage is 50µV and the integration phase is not time-constrained (see
Figure 5.3). Voltage traces for different VCM s follow each other closely except in the edge
cases VCM {0.7V , 1.2}.
48 Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier
5.1.2 Voltage Gain and Linearity
For amplifier gain analysis, we have run multiple, single data-point amplification transients
sweeping a range of input differential voltages centred around zero. These simulations are
under nominal conditions for this study: no added noise, mismatch or process variation
was included.
There are two main experiments: First, we set an integration phase run where Vin = 0
and the clk signal does not interrupt the integration process but instead lets it run its
course until both Vmida/b saturate. Thus, the crucial features of the resulting waveform
(e.g. position of peaks) are revealed. A key question we seek to answer here is whether
there is an optimum time to stop the amplification from obtaining maximum gain reliably
and, if so, when that occurs. The second experiment uses a fixed clock to explore the gain
linearity for a fixed integration period: we run multiple simulations with Vin swept from
500µV to 500µV with integration period τ= 70ns. The critical question is whether
the amplifier has a usable linear range centred around the 0Vdifferential input and, if so,
how wide it is.
The first experiment is illustrated in Figure 5.5(a). We observe that for all test inputs Vin
from 500µV to 500µV in the step of 100µV . Vmid increases linearly to a global peak
at 237ns into the integration phase and then gradually decreases to zero. At this point,
both Vmida/b have saturated and any potential difference they had is erased. The peak
occurs because as we keep integrating, the voltage at mida/b nodes eventually increases
to the point where the cascode transistors enter the triode mode. This causes the rate of
voltage accumulation on whichever Vmid node is highest to slow first, allowing the other
node to catch up (and leading to the post-peak drop in Vmid ). At this point, we are
past maximum gain and continuing the integration eventually equalises the Vmids.
Next, we note that the peak gain time is nearly perfectly aligned for all input samples; the
maximum peak time difference is only 1ps. The high quality of alignment arises because
the time at which the Vmid voltages start trioding the cascode transistors is determined
primarily by the tail current and not the differential currents. The slight discrepancy is
explained by the fact that the peak gain time is technically determined by the time at
which the first of Vmida/b reaches the point where it triodes its cascode transistor. This
has two critical engineering implications: 1) It allows us to set a universally optimal DLC
triggering time. 2) It states that the optimal trigger time is bounded by the trioding time
obtained for Vmida = min and Vmidb = max (or vice versa), in which case we have the
fastest trioding corner.
The results from the second experiment are shown in Figure 5.5(b). The differential output
voltages Vmid for τ= 70ns are plotted versus input differential voltage Vin. We notice
excellent gain linearity arising again from the minimal effect that the differential voltages
have on the behaviour of the voltages at Vmida,b. For this experiment, the differential
Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 49
140 160 180 200 220 240 260 280 300 320 340
Time(ns)
-15
-10
-5
0
5
10
15
Vmid (mV)
-500 -400 -300 -200 -100 0 100 200 300 400 500
Input Difference ( V)
-15
-10
-5
0
5
10
15
Vmid (mV)
∆Vin = -500𝜇V
∆Vin = 500𝜇V
RMES = 3.45×10-8
(a)
(b)
Figure 5.5: Simulation results of the differential gain analysis. In this simulation, inb
was set at 1Vwhile ina was swept from 1V500µV to 1V+ 500µV with in steps of
10µV . (a) Vmid throughout an intentionally excessively long integration phase. As
Vmida,b increases, the cascode transistors eventually triode causing the gain to peak and
decrease. Peak gain times occur at t= 237ns and are aligned within 1ps difference.
An indicative integration time leaving a substantial margin for error can be set to, e.g.
220ns (dashed line in (a)). (b) Output voltage difference Vmid at integration time
τ= 220ns vs input differential voltage. A linear curve excellently fits the result. The
gain is constant at approx. G= 20.
input voltage was swept based on a fixed input Vmidb = 1Vand a swept input Vinb
[1V500µV, 1V+500µV ] in the step of 10µV . Results were linearly fitted yielding a gain of
G= 20V/V (26dB) with excellent linearity throughout the range (RMSE = 3.45 ×108).
5.1.3 Input Offset Voltage
In order to access the input offset voltage, we utilise the transient simulation by multiple
transient simulations such as those seen in Figure 2.15. The measurement of offset range
is conducted in transient simulation by feeding a slow-increasing signal and a constant
reference voltage at the inputs and recording the input voltage that changes the outputs
of DLC. By tracking at what difference Vin the outputs flip value, we can obtain an
estimate for the offset. The quality of the estimate is calculated as follows: if at cycle n
we had Vouta = 0 and at cycle n+ 1 we obtained Vouta = 1.8, it means that somewhere
between Vin|nand Vin|(n+ 1) we crossed the amplifier’s offset voltage. The tracking
will be applied in ascending and descending phases, after which offset voltage will be
50 Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier
averaged. Assuming that the amplifier always decides at approximately the same relative
time in each cycle, the interval is fixed and proportional to the total swept range over the
number of sampling cycles.
0 0.5 1 1.5 2
1.0999
1.09995
1.1
1.10005
1.1001
0
0.5
1
1.5
2
Voltage (V)
0 0.5 1 1.5 2
0 0.5 1 1.5 2
1.0999
1.09995
1.1
1.10005
1.1001
0
0.5
1
1.5
2
Voltage (V)
0 0.5 1 1.5 2
Time (ms)
+500μV
V(ina) V(inb)
V(outa) V(outb)
(b)
(a)
-500μV
1.1V
Figure 5.6: Pre-amplifier basic functionality test: Input A (ina) is slowly swept between
[1V500µV , 1V+ 500µV ] over 2ms while the pre-amplifier is carrying out a conversion
every 1µs to detect the relationship between inputs A and B. Input B (inb) remains
stable at 1Vthroughout. In this test, the amplifier was balanced (R1 = R2).
In our case, we run 2000 cycles (1µs/cycle for a total duration of 2ms) and sweep the
input across a range of 1000µV (ascending from 500µV to 500µV and descending from
500µV to 500µV ), yielding an offset estimate resolution of 1µV . After setting up the
transient simulation, we apply this to Monte Carlo simulation that includes both process
and mismatch to obtain the input offset voltage. The result of the Monte Carlo simulation
is presented in Figure 5.7 and the offset voltage is summarised in Table 5.2. The 99.7%
(3σ) of the input offset voltage is within around 200µV . Therefore, we can seek the
tuneable range covering both neural spikes and circuit inherent offset which requires the
input differential is greater than 700µV .
Table 5.2: Results of a normal distribution from Monte Carlo simulation.
Parameters Offset (µV ) Parameters Offset (µV )
Mean 6.58 2σ138.92
StdDev -0.93 2σ-125.76
σ72.75 3σ205.09
σ59.60 3σ191.39
Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 51
-240 -200 -160 -120 -80 -40 0 40 80 120 160 200 240
Offset Voltage ( V)
0
5
10
15
20
25
30
35
40
45
No. of Samples
-
σσ2σ3σ
-2σ
-3σ
𝜇
Figure 5.7: Histogram of Monte Carlo simulation with 200 samples of the input offset
voltage.
5.1.4 Tunable Range and Sensitivity
To obtain the tuneable range and sensitivity of implanted memristive devices, multiple
transient simulations such as those seen in Figure 5.6 can be repeated while sweeping
both memristor device resistive states (R1 and R2).
Table 5.3 shows the offset voltage as a function of R1, R2. From there, we observe: 1) The
overall trimming range for this particular design can cover the neural spike amplitude.
2) As expected, the maximum induced offset occurs at the maximum R1, R2 imbalance
corners. 3) The offset sensitivity is close to 2µV/k for any combination of R1, R2. 4)
The table is almost symmetric (as expected). The slight asymmetry indicates that the
common mode voltage influences the offset voltage. This effect will be the subject of a
dedicated study. Finally, the quoted offsets were checked and are the same both on the
upward and the downward slopes, indicating no history dependence.
5.2 PSS Analysis
After transient simulation, we induce periodic steady-state analysis to simulate the fre-
quency response of the proposed front-end. The front-end is operated under a discrete
mode controlled by clock signals. The nodes of mida/b keep increasing during the inte-
gration, which features the amplifier dynamic operation points. Moreover, the DLC is
dynamic that does not have a static operation point. The conventional DC/AC analysis
is not applicable to the proposed front-end. We can use the periodic steady-state anal-
ysis to complete the performance metric in this case. The PSS analysis linearises the
periodic time-varying operating points and executes frequency conversion, which can be
52 Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier
Table 5.3: The offset voltage of pre-amplifier vs memristor device resistive state is
quoted at 1µV resolution.
R1\R2 1k 10k 20k 40k 60k 100k 140k 190k 240k 250k
1k 0 50 95 160 213 319 406 512 646 701
10k -53 0 48 117 172 266 353 460 519 633
20k -101 -48 0 70 125 219 306 412 542 601
40k -169 -117 -70 0 55 150 237 343 472 501
60k -224 -172 -125 -55 0 95 182 288 419 458
100k -319 -267 -219 -150 -95 0 87 192 323 362
140k -406 -353 -306 -237 -182 -87 0 105 229 265
190k -512 -460 -413 -343 -288 -192 -105 0 128 201
240k -646 -592 -543 -472 -4109 -323 -230 -128 0 101
250k -701 -633 -601 -512 -458 -362 -265 -201 -101 0
used for the periodic small-signal analysis [95]. It can assist us in conducting Fourier
transform based on periodic operation points and provide more visual results in the fre-
quency domain. The frequency data from PSS is utilised in periodic AC (PAC), periodic
transfer function (PXF) and periodic noise (pnoise) that takes the noise folding effects
into account. Thus, periodic analysis can be applied to this dynamic circuit to obtain the
frequency-related performance such as voltage gain, bandwidth, common-mode rejection
ratio (CMRR), power supply rejection ratio (PSRR) and input-referred noise.
The pre-amplifier is dynamic, and its significant state is at 220ns in Fig. 5.2 when
mida&midb trigger the DLC and the outputs of DLC are in metastable point (∆Vout =
5mV ), especially for measuring the performance of voltage gain and noise. Transient
simulation needs to be conducted first to locate the time when the output metastable
point appears. Then, the sample and hold module can be applied to collect the status from
both mida/b and outa/b for PSS analysis. So, the simulator can present the performance
directly, such as 1) mid-band voltage gain and bandwidth of OTA, 2) transfer functions
for both CMRR and PSRR, and 3) noise performance at a significant state.
5.2.1 Open Loop Voltage Gain and Bandwidth
From the results from Figure 5.8(a), the 3dB bandwidth of this system is [8Hz, 16.3kH z]
with 26dB mid-band gain that is eligible to cover the frequency of spikes, which is consis-
tent to the transient simulation result. The open-loop network OTA only sets the lower
cut-off frequency and maintains the initial pre-amplifiers mid-band gain and higher cut-
off frequency. Meanwhile, it rejects the DC offset and allows the memristive devices to
compensate only for the offset arising from PVT variations and mismatch. Compared
to the previous work, the transition to a triggering scheme has reduced the gain in this
implementation. However, this self-triggering regime is more robust on clock jitter and
Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 53
period variations.
100103105
Frequency (Hz)
101
102
Vmid(dB)
100103105
Frequency (Hz)
10-2
10-1
100
101
102
103
Output noise nV/sqrt(Hz)
26dB
816k
Figure 5.8: OTA gain and input noise as a function of frequency. (a) The open-loop
network OTA presents bandpass filtering with a mid-band gain of 26 dB. (b) Output
noise of the detection cycle from (periodic) noise simulation.
5.2.2 Noise Performance and Detection Accuracy
Before analysing pnoise and transient noise, we look into the kickback noise on two voltage
nodes: 1) the input of the core amplifier (ina/b) and 2) the input of the DLC (mida/b).
The kickback noise on the ina/b appears when turning on/off the core amplifier. When
turning on the amplifier, bias current charges the parasitic capacitors of input transistors.
The kickback noise can be discharged in 150ns reset phase (Fig. 5.2), and does not
influence the detection phase. Moreover, the charge distribution induces kickback noise
on the input nodes when we turn off the amplifier, which does not impact the operation
of spike detection. In addition, the voltage changes in latch nodes are coupled to the
input of DLC (mida/b) via the parasitic capacitors of transistors, which happens when
resetting the DLC. Under the same regime, the kickback noise is settled in the reset phase.
Thus, we proved that the 150ns reset phase eliminates the impact of kickback noise on
the detection result.
pnoise analysis The pnoise simulation of the discrete-time system is different from the
continuous-time system. In a discrete-time system, the noise from all phases will be
integrated into the power spectral density (PSD), considering the noise folding effects.
In contrast, the noise of a continuous-time system is only based on one static operation
point within the frequency ranges. To measure the noise, we refer to [96] which illustrates
the principle of measuring the comparator-based circuit where to capture the noise jitter
(shown in Figure 5.9). As shown in Figure 5.9, ‘periodic noise jitter analysis’ captures
the differential output level (internal latching nodes) of the DLC at the threshold point.
In this case, we set the threshold point as 5mV where the comparator is still in unlatch
stage.
54 Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier
t (s)
ΔVout (V)
Threshold
timing jitter distribution
voltage noise
distribution
Figure 5.9: Principle of periodic noise jitter analysis at the threshold point.
This result can reflect the noise contribution from the OTA and the input stage of the
DLC. Output noise is presented in Figure 5.8(b) and the integrated noise from 1Hz to
10kHz is 9.84 ×104V. Since the noise contribution is captured at the Vout = 5mV ,
the input-referred noise is equal to the integrated output noise divided by voltage gain.
And it yields 9.84µVrms input-referred noise. The low fcindicates that the flicker noise
does not affect the system since it is out of the operating frequency range.
Detection Accuracy Transient noise simulation for testing the detection accuracy was
also conducted to complement noise analysis [73]. The principle is that we fit the pre-
amplifier detection accuracy to the Gaussian distribution and determine the input-referred
noise as the distribution’s standard derivation (σ). In this case, the transient noise should
be included in this simulation. Under the condition of noise within the frequency range
of [1Hz,100M Hz], we ran 100 detection cycles for different Vin from 200µV to 200µV
in the steps of 10µV respectively and obtained the accuracy in Figure 5.10. The accuracy
is counted when the system outputs the correct answer as expected by the input signal.
After obtaining the discrete data of the detection accuracy versus the differential voltage,
we can fit these discrete data into Gaussian distribution to obtain the input-referred noise
as the standard derivation (σ).
The simulation was taken under two circumstances: R1 = R2 = 10k and R1 = R2 =
250k to test the degree that increasing the RS induces noise and reduces the accuracy.
The results show that the 250k RS only reduce the maximum of 2% accuracy under
the low differential input voltage. It indicates that increased RS to 250k does not
materially impact detection accuracy. In this case, the input-referred noise is 54.36µVrms
from the Gaussian fitting which is much larger than the pnoise result. It is because
the transient result includes the latch stage of the DLC. Under the application scenario
of detecting a spike, we can increase the sampling frequency to improve the detection
accuracy. Considering that one detection cycle is 250ns, the system can reach maximum
4MHz sampling rate.
Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 55
-200 -150 -100 -50 0 50 100 150 200
V
in ( V)
50
60
70
80
90
100
Detection Accuracy (%)
R1 = R2 = 10kΩ R1 = R2 = 250kΩ
𝜎= 54.36 µV 𝜎= 55.06 µV
Figure 5.10: Detection accuracy of different Vin from transient noise simulation under
the conditions of R1 = R2 = 10k (blue) and R1 = R2 = 250k (orange). The Gaussian
fittings and input-referred noise/standard derivation (σ) are given.
5.2.3 CMRR and PSRR
Both CMRR and PSRR are measured in PXF analysis to complement the performance
metric. In PXF analysis, we sample the output nodes of DLC when Vout = 5mV and
select the response to the common mode voltage source and power supply voltage, respec-
tively. Both CMRR and PSRR in this design are above 100dB within the bandwidth,
presented in Figure 5.11. The performance is summarised in Table 5.4.
(a) (b)
100101102103104105
Frequency (Hz)
60
70
80
90
100
110
CMRR (dB)
100101102103104105
Frequency (Hz)
80
90
100
110
120
130
PSRR (dB)
103dB
119dB
Figure 5.11: CMRR and PSRR of the open-loop network pre-amplifier.
56 Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier
Table 5.4: Performance metrics of the discrete-mode threshold detection system.
Performance Value
CMOS Tech. 0.18µm
Power Supply 1.8V
Bias Current 3µA
Duration of one detection (i)150 + (ii)70+(iii)30=250ns
Gain (midband) 26dB
Bandwidth 8Hz 16kHz
CMRR 103dB
PSRR 119dB
InputNoise 9.84µVrms @ input stage
54µV with latching
Memristor tuning range 250k covers ±700µV
Memristor tuning sensivity 2µV/k
Energy (250ns) 1.44pJ
Power (250ns) 5.77µW
5.3 Power Consumption
The power consumption has to be assessed for all operating phases of the pre-amplifier.
The most power-hungry phase is the reset phase since it is the only one with a DC path
between the power supplies. For this reason, the reset phase should be kept as short as
possible. However, during the reset phase, the core amplifier reaches a steady state at
all nodes so that the integrating phase can commence without any history dependence,
i.e. influence from or ‘memory of its previous inputs. Next, the cost associated with
the integration and digitisation phases can be split into two main components. First, the
integration cost is equal to charging the core amplifier’s capacitors from GND to their
equilibrium level, where the integration self-terminates (1.26Vin our case. Noted how
this integration cost currently spans both integration and digitisation phases because we
do not stop the integration once we trigger the DLCs). Second, the comparison cost equals
the energy needed to operate the DLC. Finally, during the ‘off’ phase, power dissipation
is mainly down to leakages.
In this design, the tail current in the operation mode (top = 250ns) is Iop = 3.2µA with
a supply voltage of VDD = 1.8V. Thus, the power consumption during operation is
Pop =Iop ×VDD = 5.77µW (5.1)
However, we utilised the ‘start-stop’ operation scheme, which has a long duration of off
mode. If we operate the amplifier at sampling rates of 20kHz that is twice the neural
spikes (one detection is in T= 50µs), the total energy consumption is
Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 57
E=Pop ×top +Ioff ×VDD ×tof f = 1.444pJ (5.2)
where the leakage current is Iof f = 17pA in the off mode (tof f =Ttop = 49.75µs).
Thus, the average power consumption is
Pavg =E
T= 28.88nW (5.3)
In summary, this ‘start-stop’ operation scheme allows the system to achieve 30nW aver-
age power consumption per channel, whose power consumption is two orders of magnitude
lower than the front-ends reviewed in Table 2.4.
5.4 Clock Skew and Jitter
This section discusses the clock skew and jitter effects on the front-end operation. It is
a dynamic system with a ‘start-stop’ scheme, where the clocking signals control the four
phases: clk ana,clk anabar,clk rst, and clk. Due to the DLC being triggered by clk
signal, we refer to this ‘forced-trigger’ operation, which is compared to the ‘self-trigger’
operation in section 6.7. The clocking signals are presented in Figure 5.12 with jitter
effects presented in dash lines.
0 50 100 150 200 250 300
0
1
2
Voltage (V)
0 50 100 150 200 250 300
0
1
2
Voltage (V)
0 50 100 150 200 250 300
0
1
2
Voltage (V)
0 50 100 150 200 250 300
Time (ns)
0
1
2
Voltage (V)
V(clk_ana)
V(clk_anabar)
V(clk_rst)
V(clk)
(i) (ii) (iii) (iv)
(iv)
Figure 5.12: Clocking of the AC-coupled front-end with the jitter effect presented
in dash lines. The timing diagram presents four phases: i) reset, ii) integration, iii)
digitisation and iv) off phases.
58 Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier
Clock Skew means different arrival times of the clocking signal to the different channels.
The propagation delay among channels can be neglected in this circuit for the following
reasons. The operation mode only occupies 250ns in every detection cycle which lasts
for 50µs. The impact from clock skew that shifts the operation mode within the long
detection cycle 50µs can be neglected in this case.
Clock Jitter means the clock signal varies at the ideal edge that the duty cycle of the
clock signals is different to the ideal value, which is presented as dash lines in Figure 5.12.
In this case, we divide the discussion into two parts: i) grey lines and ii) orange lines in
Figure 5.12. As for grey lines, clock jitters affect the reset, digitisation and off phases,
where we have set sufficient time margin that can tolerate 15ns jitters, which is explained
below. All clocking signals control the reset phase while clk rst determines the duration of
this phase mainly. The reset phase is designed for discharging the load capacitors, which
can be completed in 120ns and has a 30ns time margin. The worst acceptable case is
that clk rst shortens the reset phase to 120ns, which indicates the front-end can tolerate
15ns time jitters with the consideration of the reset phase. As for the digitisation phase,
it is controlled by clk ana,clk anabar and clk signals. In this case, mida/b is fed into the
DLC, and digitisation occurs in picoseconds when clk activates and activates the DLC.
The worst acceptable scenario is to shorten the duration of this phase to picoseconds.
Thus, we can obtain that the digitisation phase can tolerate 15ns jitter.
0 50 100 150 200 250 300 350 400 450 500
Time (ns)
0
0.5
1
Vmid (mV)
140 160 180 200 220 240 260 280 300 320 340
Time(ns)
-15
-10
-5
0
5
10
15
Vmid (mV)
-500 -400 -300 -200 -100 0 100 200 300 400 500
Input Difference ( V)
-15
-10
-5
0
5
10
15
Vmid (mV)
∆Vin = -500𝜇V
∆Vin = 500𝜇V
(a)
(b)
Figure 5.13: Simulation results of (a) differential output voltage of OTA (∆Vmid) and
(b) output voltage (Vmid). The clk signal occurs at 220ns ideally and samples the output
of OTA, whilst the tolerant variation range is within orange dash lines.
Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 59
The most vulnerable part is the integration phase, where the voltage gain depends on
the time (shown in orange dash line in Figure 5.12 and Figure 5.13). It can be observed
in Figure 5.13 that clk are supposed to trigger the DLC before 237ns in order to keep
all transistors in OTA in the saturation region (details are given in section 5.1.2). The
lower boundary of clk can be set at 180ns to guarantee that the output of OTA Vmid
can be integrated to 0.5V to meet the threshold voltage DLC. With the consideration of
symmetry, we determine the tolerant variation range is between 203ns and 237ns with
the ideal triggering time at 220ns. The integration phase lasts for 70ns ideally and its
acceptable range is in ±17ns. Since the integration phase is controlled by clk rst and clk
with two clock edges, the tolerant jitter for each is 8.5ns in this case. In summary, the
time jitters of clk ana and clk anabar are acceptable within 15ns. The jitters of clk rst
and clk are supposed to be confined to 8.5ns.
5.5 Simulation with Spike Train
In this section, we simulate the proposed front-end with a modelled ‘spike train’ with a
maximum amplitude of 500µV and the frequency is 10kHz. The spike train generation
method is in section 5.5.1, followed by the simulation setup and analysis in section 5.5.2.
5.5.1 Spike Train Generation
The spike train generation schemas are presented in Figure 5.14, where the spike train
outputs at the port ‘AP’. In this circuit, we utilise the periodic pulse and switch to
sample the sine wave to model the neural spikes train. The simulation results are present
in Figure 5.15.
V2V1
R
switch
AP
amp = 500uV
freq = 10kHz
Vop = 0
Vcl = 1.8V
V1 = 0
V2 = 1.8V
width = 40us
period =80us
Figure 5.14: Schematic of spike train generation, where the output is at the port ‘AP’.
60 Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
-500
0
500
Voltage ( V)
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0
1
2
Voltage (V)
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (ms)
-500
0
500
Voltage ( V)
(a)
(b)
(c)
Figure 5.15: The timing diagram of spike train generation. (a) is sine wave with the
amplitude of 500µV and frequency of 10kHz. (b) is the periodic pulse, and (c) is the
generated spike train.
5.5.2 Spike Train Simulation
After generating the spike train, we feed it into the electrode model (in Figure 2.5 and
Table 2.3) to append 15.77mV DC offset. Then, the stimulus will be fed into the circuit
under test to model the practical detection. The diagram of testbench is presented in
Figure 5.16.
Spike Generator Electrode Model
Front-End
ina inb
AP AP_OS
Figure 5.16: The block diagram presents the module and connection of the testbench.
The simulation result of the input signals is given in Figure 5.17. It can be seen in
Figure 5.17(b)-(e) that the glitches occur when the OTA is turned on and off. The
glitches represent the kickback noise that occurs when voltage variation in the circuit is
coupled to the input, thereby perturbing the input signals. However, there exists a reset
phase to discharge the load capacitors and initialise the OTA for detection. Consequently,
the kickback noise can be disregarded in this design. Figure 5.17 (b), (c) and (e) describe
Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 61
the two processes: i) the spike is appended with 15.77mV DC offset and ii) the DC offset
is rejected by the highpass module at the input of OTA (in Figure 5.1).
0 2 4 6 8 10 12 14 16 18 20
0
1
2
Voltage ( V)
0 2 4 6 8 10 12 14 16 18 20
-500
0
500
Voltage ( V)
0 2 4 6 8 10 12 14 16 18 20
15
16
17
Voltage (mV)
0 2 4 6 8 10 12 14 16 18 20
1
1.005
1.01
Voltage (V)
0 2 4 6 8 10 12 14 16 18 20
Time ( s)
-500
0
500
Voltage ( V)
V(clk_ana)
V(AP)
V(AP_OS)
V(ina)V(inb)
V(ina -inb)
(a)
(b)
(c)
(d)
(e)
Figure 5.17: The timing diagram of the control signal and spikes. (a) clk ana controls
the on and off status of the OTA. (b) Spike train (AP ) from spike generator. (c) shows
the spike with 15.77mV DC offset appended. (d) presents the input voltages of OTA. (e)
is the differential input voltage of OTA.
The simulation results of spike detection are presented in Figure 5.18 within 50µs. In
this simulation, the threshold voltage can be determined by setting the memristor RSs
as R1 = 1k and R2 = 5k according to Table 5.3. In this case, we set the detection
cycle as 10µs (sampling rate as 100kHz) to demonstrate that the system can detect the
spike (above 50µV ) and output the digital signal in outb. (Noted that this simulation is
noiseless.)
The practical detection is supposed to consider the noise; thus, we conduct the tran-
sient noise simulation below to determine the detection accuracy under the noise in the
frequency range of [1Hz,100M Hz]. Figure 5.19 compares the digital outputs of spike de-
tection in the condition of (b) the noiseless simulation and (c) transient noise simulation,
respectively. The blue lines in Figure 5.19(b)-(c) are the actual signals and we highlight
the output by orange lines for more explicit demonstration. The detection is 100% ac-
62 Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier
0 5 10 15 20 25 30 35 40 45 50
0
1
2
Voltage (V)
0 5 10 15 20 25 30 35 40 45 50
-500
0
500
Voltage ( V)
0 5 10 15 20 25 30 35 40 45 50
-10
0
10
Voltage (mV)
0 5 10 15 20 25 30 35 40 45 50
Time ( s)
0
1
2
Voltage (V)
control signals
(a)
(b)
(c)
(d)
V(ina-inb)ideal spike
V(midab)
V(outa) V(outb)
Figure 5.18: Simulation result of spike detection. (a) presents control signals which
are the same as Figure 5.2. (b) presents the differential input voltage ina inb and the
ideal spike. (c) shows the differential output voltage of OTA, and (d) presents the digital
outputs.
0 50 100 150 200 250 300 350 400
-500
0
500
Voltage ( V)
0 50 100 150 200 250 300 350 400
0
1
2
Voltage (V)
0 50 100 150 200 250 300 350 400
Time ( s)
0
1
2
Voltage (V)
(a)
(b)
(c)
ideal spike
V(outb) without noise
V(outb) with transient noise
Figure 5.19: Simulation result of spike detection within 400µs. (a) presents the ideal
spike. (b) is digital output in noiseless simulation, and (c) is obtained from transient
noise simulation.
Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier 63
curate in noiseless simulation, while it reaches 80% accuracy with the consideration of
noise.
As discussed above, ‘detection accuracy’ refers to the proportion of correct results. The
‘detection accuracy’ is fixed when the circuit is designed. Furthermore, to determine
whether a spike exists, we need to induce a voting module in the back-end. We can utilise
the sampling frequency and ‘detection accuracy’ to determine whether there is a spike.
For example, if the oversampling ratio is 11, accurate detection equal to or greater than
six can be considered a successful capture of the spike. This can be accomplished by using
permutation and combination. We refer to this as ‘spike detection accuracy’. Thus, in
Figure 5.19, the ‘detection accuracy’ is 80% and the ‘spike detection accuracy’ is 87.5%
(correct spike detection: seven out of eight, presented in Figure 5.20(a)). In this case, the
detection cycle is 10µs, and the average power consumption of the front-end is 144nW .
Moreover, when we increase the sampling frequency to 200kH z (each detection cycle is
5µs), the spike detection accuracy reaches 100%, and the average power consumption is
only 288nW .
0 50 100 150 200 250 300 350 400
-500
0
500
Voltage ( V)
0 50 100 150 200 250 300 350 400
0
1
2
Voltage (V)
0 50 100 150 200 250 300 350 400
Time ( s)
0
1
2
Voltage (V)
ideal spike
0 50 100 150 200 250 300 350 400
-500
0
500
Voltage ( V)
0 50 100 150 200 250 300 350 400
0
1
2
Voltage (V)
0 50 100 150 200 250 300 350 400
Time ( s)
0
1
2
Voltage (V)
V(outb) with transient noise
0 50 100 150 200 250 300 350 400
-600
-400
-200
0
200
400
600
Voltage ( V)
0 50 100 150 200 250 300 350 400
Time ( s)
0
0.5
1
1.5
2
Voltage (V)
(a)
(b)
V(outb) with transient noise
ideal spike train
(a) (b)
Figure 5.20: The timing diagram is to present the spike detection accuracy. The
detection cycle is 10µs in (a) and 5µs in (b), respectively.
5.6 Discussion
From the analysis and simulation of the integrating amplifier, we highlight some key
conclusions:
First, the performance improvement of the integrating amplifier over more traditional
designs relies on the integration process, which enhances the gain and decreases the effec-
tive bandwidth (helping reduce noise in the process). To visualise this, let us consider an
integrating amplifier using the same tail current as a standard OTA first stage. During
integration, the power dissipation is the same, but the gain and bandwidth differ. In this
sense, the design represents a trade-off between gain and bandwidth without changing
power dissipation or using feedback.
64 Chapter 5 Methodology of Measuring Memristor-based AC-Coupled Pre-amplifier
Next, we note that there is a natural trade-off between tail current and integration time
while keeping the overall energy dissipation approximately constant. This is the result of
the fixed duration of the reset phase (just enough to clear any residual charge at the Vmid
nodes) and the fact that energy consumption during the integration phase only depends
on the size of the load caps and the voltage change across them during that phase. Thus,
we can design a wide range of required sampling rates or bandwidths for the same energy
budget.
The trade is not entirely free: Changing the tail current affects gain, bandwidth and noise
performance by altering the gms of all transistors involved and the integration period.
Furthermore, if using real memristor devices with non-linear IV curves, changing the tail
current also changes the static resistance of the memristor devices. Together with changes
in transistor gms, the tuneability range is also affected since it depends on the impedance
balance between memristor and transistors. Thus, whilst the integrating amplifier offers
a lot of design flexibility, the precise design trade-off space is also not trivial, much like it
is for OpAmps. This is a critical subject meriting its dedicated study.
The last design decision to highlight concerns the size of the load capacitors C. The gain
analysis shows that Cdoes not affect the gain, but it does affect the integration period
and, therefore, can be used to adjust the bandwidth if, for some reason, that cannot be
achieved by tweaking the tail current. Effectively it is a design parameter that trades
away energy for design flexibility.
In terms of operation, we note the importance of ensuring that the integrating amplifier
is cleared properly in preparation for each integrating phase to avoid the output’s history
dependence. This means that all node voltages should be equalised across the left and
right branches prior to the commencement of the sensitive integration phase. The current
design achieves this by forcefully flushing the system during the reset phase. However,
more energy-efficient approaches are under development as the rest phase represents a
substantial fraction of the energy budget.
5.7 Summary
In this chapter, we apply the proposed OTA to the open-loop network and conduct the
transient simulation and periodic steady-state analysis to complete the performance met-
ric of the pre-amplifier. This AC-coupled pre-amplifier utilises resistors and large input
capacitors to reject the DC offset. From this design, we derive the design trade-off and de-
sign consideration of this start-stop scheme. Moreover, the proposed AC-coupled front-end
is proven to remove DC offset and achieve 100% spike detection accuracy (in demonstra-
tion), only consuming 288nW average power consumption per channel. Furthermore, this
version pre-amplifier has been fabricated, and its layout design and the testing result are
presented in Chapter 7.
Chapter 6
A Memristor-based DC-Coupled
Pre-amplifier Topology
6.1 The Proposed Circuit
In this chapter, we propose a hybrid CMOS/memristor front-end that can compensate
the input DC offset up to 50mV . The block diagram of the front-end is presented in Fig-
ure 6.1(a). The electrode’s reference voltage and neural signal are fed into the memristive
OTA from nodes of inb and ina, respectively. The tissue-electrode model in Figure 6.1(b)
is referred to Jochum et al. [97]. The simplified schematic is shown in Figure 6.1(c). The
schematic of this OTA is similar to the AC-coupled circuit in Chapter 4and 5, except
for the position of memristors. The detailed schematic is in Figure 4.1 and the informa-
tion on transistor sizes is in Table 6.1. The memristors are assembled below the input
transistors to tune the branch currents in a wider range, compared to the AC-coupled
solution in Chapter 5. Note that the methodology of simulating the circuit is the same
with Chapter 5provided.
Table 6.1: Sizes of devices in the proposed architecture, where the bias current of the
core amplifier is Itail = 3µA.R3 is replaced by a diode-connected NMOS. The detailed
schematic is in Figure 4.1.
Devices W/L (µm/µm) Devices W/L (µm/µm)
M1, M2 1.2/2 R3 1/1
M3 1.2/2 M11-M14 2/0.6
M4, M5 16/1 M15,M16 1/0.6
M6, M7 4/1 M17, M18 2/0.6
M8, M9 1/0.6 M19 1/0.6
M10 2/1 C1, C2 100 fF
65
66 Chapter 6 A Memristor-based DC-Coupled Pre-amplifier Topology
RHF
RLF
CPE
RRAM
OTA DLC
outa
midb
mida
inb
ina
ina inb
clk_ana
mida midb
R1 R2
clk_anabar
clk_rst
(a)
(b) (c)
clk
outb
CL1 CL2
Figure 6.1: DC-coupled front-end. (a) Block diagram of the neural recording front-end.
(b) The electrical model of the tissue-electrode interface. (c) Simplified schematic of the
OTA assembled with memristive devices.
6.2 Transient Analysis
This chapter presents a new clocking scheme in Figure 6.2.The system is controlled by
four clocking signals: clk ana,clk anabar,clk rst and clk in Figure 6.2(a). The discrete-
mode front-end is divided into three operation phases: (i) reset, (ii) detection and (iii) off
phase. In the reset phase, the amplifier is switched on and the current branch from V DD
to the ground is connected. The 150ns reset phase helps initialise the amplifier and DLC
and discharge the load capacitors CL1,2.
In the detection phase, the system conducts signal integration and digitisation in DLC.
Once the mida/b integrate above the threshold voltage of DLC (0.5V), the DLC is trig-
gered and processes the digitisation based on the differential voltage mid =VmidbVmida
(in Figure 6.2(a)-(d)). We highlight that the ‘detection phase’ combines the ‘integration
phase’ and ‘digitisation phase’ in Figure 2.14. In the ‘integration phase’ of the initial
design, the DLC is off until Vmid is integrated to the maximum voltage and the pre-
amplifier enters the ‘detection phase’ (we named this scheme as ‘forced-trigger’). On the
other hand, the DLC is enabled and awaiting activation by the output of the core OTA
(termed ‘self-triggered’). This scheme prevents the jitters of the controlled clocking signals
from fluctuating the input signals of the DLC and causing glitches.
Chapter 6 A Memristor-based DC-Coupled Pre-amplifier Topology 67
Once the detection completes, the amplifier is turned off. The duration of APs is about
a few milliseconds, and APs occur from 10 to 120 per second [45]. It indicates that the
front-end can be in the off mode in the interval of spikes to save power consumption.
Besides, the low spike frequency provides much headroom for oversampling.
0 50 100 150 200 250
0
1
2
Voltage (V)
0 50 100 150 200 250
0.95
1
1.05
Voltage (V)
0 50 100 150 200 250
0
0.5
1
Voltage (V)
0 50 100 150 200 250
0
2
4
6
Voltage (mV)
0 50 100 150 200 250
Time (ns)
0
1
2
Voltage (V)
(i)
(ii)
(iii)
V(clk_ana) V(clk_anabar) V(clk_rst)
V(clk)
V(midb-mida)
V(outa) V(outb)
V(mida)
V(ina)
(a)
(b)
(c)
(d)
(e)
Figure 6.2: The timing diagram of one neural detection cycle (250ns). (a) clocking
scheme. (b) input voltage of the front-end from the electrode. (c) the output voltage of
the core amplifier. (d) the differential output voltage of the amplifier. (e) digital outputs
from the DLC. (In this simulation, ina = 1V+ 50µV and inb = 1V.)
6.3 PSS Analysis
The primary performance metrics of the front-end are presented in Table 6.2, where most
of the indicators are measured at the nodes of mida/b, presenting the performance of
the memristive amplifier. Moreover, the energy and power consumption is of the whole
front-end. The integration process can amplify the differential input signal from micro-
volt to milli-volt, which makes it recognised by the DLC. The gain is 21.2dB within the
bandwidth of 13.7kHz. The CMRR of the memristive amplifier reaches 98.7dB, while
the PSRR is 63dB from the PXF analysis. Since we stated that the input-referred noise
from discrete mode is not comparable to the conventional amplifier, we can only extract
68 Chapter 6 A Memristor-based DC-Coupled Pre-amplifier Topology
that the front-end can operate in accurate DC with 394nV/Hz white noise.
The energy and average power are measured for one detection cycle within 250ns. As
the neural spike is in low frequency and at a low occurrence rate, the average power
consumption also depends on the sampling rate. Therefore, if the front-end operates
under 20kHz sampling rate, the average power can be reduced to 29nW . The discrete-
mode operation with low power consumption provides lots of headroom to increase the
sampling rate and detection accuracy.
Table 6.2: Performance metric of memristor amplifier (The energy and average power
consumption are of one detection cycle with 250ns, including both memristor amplifier
and DLC).
Performance Value
Power Supply 1.8V
Bias Current 3µA
Gain 21.2dB
Bandwidth 13.7kHz
CMRR 98.7dB
PSRR 63.0dB
Input Noise 1/f corner: 21.9Hz
thermal noise: 394nV /Hz
Energy (250ns) 1.44pJ
Power (250ns) 5.77µW
6.4 Tuneable Offset Range
In this simulation, we import the memristor model by Maheshwari et al. [35]. In or-
der to measure the tuneable offset range, a testbench is designed with the resolution of
10µV . For example, if the input voltage is from zero to 1mV , we set the detection to
appear every microsecond and run the simulation for 100 cycles. Then, the digital output
outa/b will switch at some point when the Vin keeps increasing. Furthermore, the Vin
corresponding to the switch point is the offset input voltage.
The tuneable offset range is summarised in Table 6.4, where R1 = 10k and measures
the R2 in the step of 10k till the offset range reaches 50mV . The applied memristor
can reach 130k [27]. We can obtain from the table that: 1) the tuning sensitivity varies
from different R2 that the larger R2 induces lower sensitivity, where the resolution can
reach 19µV/kΩ; 2) 1% RS variation causes maximum 6.7% drift of the offset range. The
accurate RSs that compensate the offset to {1,10,50}mV are measured as well (shown in
Table 6.3).
Chapter 6 A Memristor-based DC-Coupled Pre-amplifier Topology 69
Table 6.3: The compensated DC offset versus the resistive state of memristors under
R1 = 10 kΩ.
R2 (kΩ) Offset Compensation (mV )
49 1
83.6 10
127.3 50
Table 6.4: Memristor tuneable offset range when R1 = 10k and R2 is from 10k to
130kΩ. The third column presents the tuneable range with 1% RS variation from two
memristors. The offset resolution is 10µV .
R2 (kΩ) Nominal
offset (V)
Tuneable offset with 1%
RS variation (V)
Offset variation
(%)
10 0 10µ10µ /
20 150µ140µ150µ6.7%
30 340µ330µ340µ2.9%
40 610µ600µ630µ3.3%
50 1.09m1.05m1.12m3.7%
60 2.05m1.97m2.14m4.4%
70 4.14m3.94m4.35m5.1%
80 8.08m7.68m8.48m5.0%
90 14.09m13.47m14.73m4.5%
100 21.96m21.10m22.84m4.0%
110 31.41m30.22m32.42m3.8%
120 41.82m40.51m43.14m3.2%
130 53.17m51.66m54.70m2.9%
6.5 Noise Analysis
Kickback noise appears on the inputs of both the memristive amplifier and the DLC. In
Figure 6.1(b), the kickback noise happens when switching on/off the amplifier. When the
amplifier is turned on, the bias current charges the parasitic capacitors of input transistors.
However, it does not impact the detection result because the long reset phase allows the
kickback to be settled before the detection phase. Furthermore, when we turn off the
amplifier, the kickback noise appears due to the charge distribution, which happens in
the off phase and does not influence the spike detection. From Figure 6.2(d), kickback
noise exists on mida/b due to the voltage changes in the latch nodes when resetting the
DLC. Still, it was eliminated before the detection phase.
The detection accuracy is measured using a transient noise simulation with 250 runs.
In order to estimate the noise effect, we apply resistors as memristors because there is
no mature memristor noise model. In this case, we keep R1 as 10k and set R2
{49k, 83.6k, 127.3k} to imitate the circuit status with {1m, 10m, 50m}Voffset compen-
sation. Then, Vin is swept from 10µV to 500µV in the step of 10µV . The detection
accuracy is in Figure 6.3.
70 Chapter6A Memristor-basedDC-CoupledPre-ampliierTopology
10
(μV)
Figure6.3: Detectionaccuracyoffront-endsystemvsdiferentialinputvoltages V
in
from10µV to400µV.Inthiscase,R1 =10k,R2{10,49,83.6,127.3}kinorderto
compensatetheDCofsetof
{1m,10m,50m}V.
Whentworesistorsarein10kΩandthe V
in=50µV,thedetectionaccuracyreaches
74.4%. Andtheaccuracycanreach100% whenthediferentialinputexceeds200µV.
EventhoughRSisincreasedtocompensatefor10mV ofset,theaccuracyreducesbya
maximumof9%. Thereexists moreconsiderableaccuracydownwhentheRSiscapable
ofcovering50mV ofset. Theaccuracyinthissituationreaches58.4%for50µV difer-
entialinput. However,thedetectionaccuracycanbeimprovedthroughoversampling,as
describedinthefollowingsection.
6.6 Oversampling
Theabovedetectionaccuracyisobtainedwhenthefront-endonlydetectsonceforevery
neuralspike.Increasingthesamplingrateofthefront-endandinducingmajorityvotingon
theback-endcanincreasethespikedetectionaccuracy.Forexample,iftheoversampling
ratiois11,accuratedetectionequaltoorgreaterthansixcanbeconsideredasuccessful
captureofthespike. Assumingtheoversamplingratiois
m(anoddnumber),thedetection
accuracyisAandinaccuratepossibilityis¯
A,theaccuracyis
Cm+1
2
m(A
m+1
2·
¯
A
m−1
2)+···+C
m−1
m(A
m−1·
¯
A)+C
m
m(A
m) (6.1)
whichcanbesummarisedas
Accuracy=
m
n=m +1
2
C
n
m(A
n·
¯
A
m−n
) (6.2)
ThedetectionaccuracyundertheconditionsofR2 =83.6k,127.3kand V
in=50µV
are58.4%and67.6%respectively. Moreover,theinitialvaluecanbeutilisedtoestimate
Chapter 6 A Memristor-based DC-Coupled Pre-amplifier Topology 71
the oversampling effect, which is presented in Figure 6.4. It can be estimated that when
we aim to compensate typical 10mV DC offset, the spike detection accuracy can reach
95% when the oversampling ratio is 21. In this case, we simulate the input signal with
a maximum frequency of 10kHz. Conducting 21 times indicates one detection cycle is
47µs consumes 302.4nW average power of the front-end. Moreover, under the ex-
treme situation of R2 = 127.3kΩ, to guarantee the spike detection accuracy is 95%, the
oversampling ratio needs to reach 95 and the system consumes 1.368µW .
50 100 150 200
Oversampling Ratio
60
70
80
90
100
Detection Accuracy (%)
R2 = 83.6k
R2 = 127.3k
1 21 95
Spike Detection Accuracy (%)
Figure 6.4: Spike detection accuracy under oversampling. The initial value takes from
Fig. 6.3, where the accuracy of R2 = 83.6k and R2 = 127.2k are 67.6% and 58.4%
respectively under Vin = 50µV .
6.7 Clock Skew and Jitter
In this section, we discuss the clock skew and jitter effects on the operation of the proposed
DC-coupled front-end. This front-end operation is similar to the AC-coupled pre-amplifier
operation (in section 5.4), except for the triggering of the DLC. In this design, DLC is
activated after resetting the OTA and waiting for its output to trigger. Thus, we refer to
this ‘self-trigger’ DLC. There are three phases: i) reset, ii) detection and iii) off phase,
which are controlled by the clocking signals shown in Figure 6.5. The clock skew analysis
has been conducted in section 5.4. And we focus on the jitter analysis in this section.
In the reset phase, the worst case that can be accepted is that the duration is shortened
from 150ns to 120ns when the load capacitors are fully discharged. Thus, for each edge
variation, the tolerant value is (150 120)/2 = 15ns. The detection phase contains both
integration and digitisation. The detailed operation in Figure 6.2 presents that it takes
20ns to complete the digitisation, which indicates there is no time margin for clock jitters
in this case. In order to design a front-end that is robust to clock jitters, we need to
extend the duration of the detection phase. For instance, to tolerate the 15ns clock jitter,
72 Chapter 6 A Memristor-based DC-Coupled Pre-amplifier Topology
0 50 100 150 200 250 300
0
1
2
Voltage (V)
0 50 100 150 200 250 300
0
1
2
Voltage (V)
0 50 100 150 200 250 300
0
1
2
Voltage (V)
0 50 100 150 200 250 300
Time (ns)
0
1
2
Voltage (V)
V(clk_ana)
V(clk_anabar)
V(clk_rst)
V(clk)
(i) (ii) (iii)(iii)
Figure 6.5: Clocking of the AC-coupled front-end with the jitter effect presented in
dash line. The timing diagram presents three phases: i) reset, ii) detection and iii) off
phases.
we need to extend the detection phase from 100ns to 130ns. The operation time will
be extended and increase power consumption. For example, if the front-end operates at
the sampling frequency of 20kHz, the power consumption increases from 29nW to 33nW
per channel. It indicates a design trade-off between robustness to clock jitter and power
consumption.
6.8 Spike Train Detection
In this section, we conduct the practical detection of spike train with the same simulation
setup in section 5.5. The detailed simulation within 50µs is given in Figure 6.6, followed
by the spike detection accuracy analysis.
The simulation results are present in Figure 6.6. (a) presents the ideal control signals
which are the same as Figure 6.2. (b) presents the spike with an amplitude up to 500µV .
There are glitches at the edge of the operation mode and off mode represent the kickback
noise. (c) presents the differential input signals of OTA and it can be seen that the
voltage of ina is 15mV larger than the inb (reference). The OTA is fed with tens of DC
offset and utilises the memristor to compensate. In this case, we set the R1 = 10k and
R2 = 91.2k to compensate the DC offset (15.77mV ) and detection threshold voltage
(50µV ).
This circuit operates with a ‘self-trigger’ DLC, which means the DLC is activated after
the reset phase. After that, the outputs of OTA increase gradually through integration
Chapter 6 A Memristor-based DC-Coupled Pre-amplifier Topology 73
and trigger the DLC once the voltage reaches the DLC threshold. The digitisation time,
in this case, is longer than the ‘forced-trigger’ scheme due to the low input voltage of DLC.
Thus, the long settling time can be seen in Figure 6.6(e) compared to the ‘forced-trigger’
scheme in Figure 5.18(d).
0 5 10 15 20 25 30 35 40 45 50
0
1
2
Voltage (V)
0 5 10 15 20 25 30 35 40 45 50
-500
0
500
Voltage ( V)
0 5 10 15 20 25 30 35 40 45 50
1
1.01
1.02
Voltage (V)
0 5 10 15 20 25 30 35 40 45 50
-10
0
10
Voltage (mV)
0 5 10 15 20 25 30 35 40 45 50
Time ( s)
0
1
2
Voltage (V)
(a)
(b)
(c)
(d)
(e)
control signals
V(AP)
V(ina)V(inb)
V(midab)
V(outa) V(outb)
Figure 6.6: Simulation result of spike detection. (a) presents control signals which are
the same as Figure 6.2. (b) is the spike train generated in Figure 5.16. (c) presents the
input voltages of OTA. (d) shows the differential output voltage of OTA, and (e) presents
the digital outputs.
The detection results with the sampling frequency of 100kHz and 200kHz are depicted
in Figure 6.7. The simulation is performed in the presence of [1Hz, 100M H z] transient
noise. Figure 6.7 demonstrates that the spike detection accuracy can reach 87.5% even
though the detection accuracy is 72.5%. It indicates that the user can detect the presence
of a spike with an accuracy of 87.5%. In addition, the spike detection accuracy can
reach 100% with a sampling frequency of 200kHz. According to the clock jitter analysis
in section 6.7, we can extend the operation mode from 250ns to 280ns to enhance the
robustness of clock jitter at the cost of 33nW per channel. Furthermore, if this system is
utilised at a sampling rate of 200kHz, the average power consumption is 323nW .
74 Chapter 6 A Memristor-based DC-Coupled Pre-amplifier Topology
0 50 100 150 200 250 300 350 400
-500
0
500
Voltage ( V)
0 50 100 150 200 250 300 350 400
0
1
2
Voltage (V)
0 50 100 150 200 250 300 350 400
Time ( s)
0
1
2
Voltage (V)
ideal spike
V(outb) without noise
V(outb) with transient noise
0 50 100 150 200 250 300 350 400
-500
0
500
Voltage ( V)
0 50 100 150 200 250 300 350 400
0
1
2
Voltage (V)
0 50 100 150 200 250 300 350 400
Time ( s)
0
1
2
Voltage (V)
ideal spike
V(outb) without noise
V(outb) with transient noise
(a) (b)
Figure 6.7: The timing diagram is to present the spike detection accuracy. The detec-
tion cycle is 10µs in (a) and 5µs in (b), respectively.
6.9 Summary
Memristive devices adjust the offset voltage to eliminate the need for large capacitors and
resistors at the input of a DC-coupled pre-amplifier to reject DC offset. Even though
the primary performance of a memristive amplifier is inferior to that of a conventional
state-of-the-art neural amplifier, it is adequate for use as an integrator followed by a
DLC. Simulations of transient noise have been used to test stability and robustness. The
results indicate that the front-end can compensate for large DC offsets of up to 50mV .
From spike train simulation, it can be derived that the DC-coupled front-end can achieve
100% spike detection accuracy (in demonstration) at the cost of 323nW average power
consumption.
Chapter 7
Layout Design and Chip Testing
This chapter focuses on the layout design of single channel pre-amplifier and chip test
after tape-out. The details and techniques of layout design with the program unit of the
memristive device are provided, followed by the post-layout simulation. In this work, we
used a commercially available 180nm CMOS technology with V DD = 1.8V. Further-
more, the preparation for the chip test, including test strategy and test board design, is
described. We conclude with a demonstration of the results of chip tests and the insights
gained from this experience.
7.1 Layout Design
The pre-amplifier consists of both analogue and digital parts, whereas the analogue layout
design requires high matching and low noise. Firstly, the 1.8Vpower supply of analogue
and digital modules needs to be separated, and the programming unit is supplied by 5V
voltage source. The most sensitive part of the circuit is the differential input pair and
cascode pair; thus, the bulks are connected to the source of the input pair (in Figure 7.1(a))
with the protection of two guarding rings. There are double protections: i) the guarding
ring surrounds the input pair and the cascode pair, and ii) the others isolate these two
elements from other components to prevent the sensitive circuit from interfering with
other modules (in Figure 7.1(b)).
In addition, we induced interdigitation and common centroid techniques in layout design
to access high matching. The sizes of each component have been shown in Table 7.1, where
the width of the differential input pair (60µm) is three times larger than the cascode pair
(20µm). Thus, we divided the input pair into three parts and distributed them according
to the common centroid in Figure 7.1(b).
The programming circuit in Figure 7.2(a) consists of a memristor and two transmission
gates with the combination of NMOS and PMOS. The transmission gate is utilised to
75
76 Chapter 7 Layout Design and Chip Testing
Table 7.1: Sizes of devices in the proposed architecture, where the bias current of core
amplifier is Itail = 3µA.R3 is replaced by a diode connected NMOS.
Devices W/L (µm/µm) Devices W/L (µm/µm)
M1, M2 3/3 R3 1/1
M3 3/3 M11-M14 2/0.6
M4, M5 60/1 M15,M16 1/0.6
M6, M7 20/1 M17, M18 2/0.6
M8, M9 1/0.6 M19 1/0.6
M10 2/1 C1, C2 100 fF
M4 M5
M6 M7
ina inb
clk_anabar clk_anabar
(a) (b)
Figure 7.1: Differential input and cascode pair in both (a) schematic and (b) layout
view. The bulks of the above components are connected to the source of M4&M5.
isolate the pre-amplifier (1.8Vsupply voltage) with the high-voltage control signal. In this
case, the terminals A1&A2 connect to the pre-amplifier, while terminals B1&B2 directly
connect to specified pads so that the designated signals can program the memristive
device. Since the programming signal is up to 5V, the pass-transistors are applied to
shield the core circuit from damage by high voltage. In this user-isolation switch, when
EN MEM = 0 and EN BAR = 1, the transmission gates are off and the memristor
can be programmed by the signal from the pad, referring back to Chapter 3. When
EN MEM = 1 and EN BAR = 0, the transmission gates are on so that the memristor is
in the application status. The layout of the switch is shown in Figure 7.2(b).
The ESD (electro-static discharge) cell is utilised for analogue and digital inputs/outputs,
except for memristor programming input. For the input pin, the ESD cell works as a
protection network that makes it operate under passive conditions. For the output pin,
the protection element is supposed to minimise the performance and voltage degradation
of the I/O circuit. The schematic of the ESD cell referring to the 180nm has been shown
in Figure 7.3(a), and the layout following the design instruction is in Figure 7.3(b). An
ESD cell occupies the area of 16.120µm ×30.210µm.
After accessing the essential components of the layout, we provide the layout of the core
circuit in Figure 7.4. In this layout, we included 21 pins which are presented in Table 7.2,
where we separated the analogue power (1.8V) with memristor program unit (5V) and
the digital (1.8V) ones so that we can prevent the sensitive analogue module will be
Chapter 7 Layout Design and Chip Testing 77
N1
N2
P1
P2
A1
MEM
A2
ENMEM ENBAR
B1
B2
(a)
P1 P2
N1 N2
MEM
(b)
Figure 7.2: Schematic and layout of the programming circuit of the memristive device.
(a) Terminals A and B can be the user terminal and programming terminal or in reverse.
The status of memristive in the application or programming is controlled by two signals,
EN MEM and E N BAR. In layout (b), the memristive device is in the middle and
connects to PMOS (upper device) and NMOS (lower device).
R1
PMOS
NMOS
Vdd
PadIn
PadOut
(a) (b)
Figure 7.3: The schematic (a) and layout (b) of ESD Cell based on TSMC180nm
technology.
interfered by the digital one. In addition, it can be seen that the guarding rings are
utilised to isolate modules in terms of the input and cascode pair, the current generation
unit, the programming circuit, the digital DLC circuit and the substantial load capacitors.
The area of the pre-amplifier is 71.280µm ×32.255µm.
The complete layout, including the pads, is presented in Figure 7.5. In this tape-out, we
fit the pre-amplifier into 26 in-line pads, whose total area is 2080.000µm ×283.670µm. In
this design, we utilise the pads provided by the foundry for the power, analogue and digital
signals and also use in-house electrodes directly connected to the memristive devices. It is
because that programming memristor requires a higher current compared to other pins.
78 Chapter 7 Layout Design and Chip Testing
Current Mirror
M4_1
M4_2
M4_3
M5_1
M5_2
M5_3
M6
M7
M8, 9
R1
R2
C1
C2
DLC
Figure 7.4: Layout of the core pre-amplifier circuit with the input and output on the
boundary to connect the ESD cell or the pad directly.
Figure 7.5: Complete layout design of the pre-amplifier. This design has two types of
pads: i) the commercial pad from the foundry and ii) the in-house pad for programming
the memristive devices.
Table 7.2: The pins of the pre-amplifier with the classification of type. There are
some abbreviations: ana in/out represents analogue input/output signal, ana in HV
means high voltage analogue input which is 5Vwith the supply of VDD MEM. Be-
sides, dig in/out is the digital input/output signal. pwr x, where x= 1.8/5 is the supply
voltage.
No. Pin Name Type No. Pin Name Type
1 GND AMPDLC pwr 12 clk rst dig in
2 clk anabar dig in 13 clk dig in
3 inb ana in 14 outb dig out
4 ina ana in 15 outa dig out
5 VDD AMP pwr 1.8 16 VDD DLC pwr 1.8
6 clk ana dig in 17 ENMEM 1 dig in
7 A2 2 ana in HV 18 ENBAR 1 dig in
8 VDD MEM pwr 5 19 A1 1 ana in HV
9 A1 2 ana in HV 20 A2 1 ana in HV
10 ENBAR 2 dig in 21 GND MEM power
11 ENMEM 2 dig in
Chapter 7 Layout Design and Chip Testing 79
7.2 Post-Layout Simulation
After designing the layout, we extract the layout for the post-layout simulation and make
the performance comparison between the pre-layout and post-layout pre-amplifier. For
the post-layout circuit, the first thing is to determine the inherent offset and utilise the
memristive device to compensate the offset voltage to zero. In order to obtain the inherent
offset voltage, we apply the input voltage ramp in the range of [200µV ,200µV ] into the
circuit and capture the digital output. The result presents in Figure 7.6. It presents that
the inherent offset voltage is 120µV and we tune the RS of the memristor to compensate
for the offset voltage by R1 = 1k and R2 = 23.8kΩ.
s
(a)
(b)
V(outa) V(outb)
1V
-200µV
+200µV
V(ina) V(inb)
Figure 7.6: The input voltage ramp is applied to the circuit for the post-layout circuit
to determine the inherent offset voltage. The figure shows that the inherent offset voltage
is Vos =Vina Vinb = 120µV .
After calibrating the inherent offset, we run the transient simulation on this circuit in
Figure 7.7. It can be seen from Figure 7.7(b)-(c) that the output of the core amplifier can
be settled down within phase (ii). Moreover, the afterwards simulation can be based on
this transient simulation setting.
After calibrating the circuit, we apply the same simulation methodology to this circuit
to obtain a full performance metric. The comparison table between pre-layout and post-
layout results presents in Table 7.3.
80 Chapter 7 Layout Design and Chip Testing
(d)
(c)
(b)
(a)
(i) (ii) (iii)(iv)
V(clk_ana)V(clk_anabar)V(clk_rst)V(clk)
V(midb)
V(midab)
V(outa) V(outb)
(iv)
Figure 7.7: Timing diagram for post-layout simulation.
7.3 Strategy of Chip Testing
In this stage, we can access the function testing in terms of effective gain, input range,
PSRR and CMRR, noise, tunable range and sensitivity by feeding continuous signals. In
this stage, we only have access to the functional check of the signal channel pre-amplifier.
7.3.1 Programming Memristive Device
Before utilising the memristor, we need to initialise the device by electroforming to a us-
able resistance range (approximate 25k to 200kΩ, depending on the stack) [27]. Shown
in Figure 7.2(b), the terminals with pass transistors are connected to the core ampli-
fier. Thus, electroforming must be conducted by isolating the memristor with the core
amplifier. The operation steps will be shown as follows.
1. The pass transistor needs to be turned off to ensure that the core amplifier has been
Chapter 7 Layout Design and Chip Testing 81
Table 7.3: Performance metrics comparison between pre-layout and post-layout simu-
lation.
Performance Pre-layout Post-layout
CMOS Tech. 0.18µm
Power Supply 1.8V
Bias Current 3µA
Duration of one
detection
(i)150 + (ii)70+(iii)30=250ns
Gain (midband) 26.0dB 23.2dB
Bandwidth 8Hz 16kHz 3H z 17kHz
CMRR 103.0dB 95.2dB
PSRR 119.0dB 78.6dB
InputNoise 9.84µVrms @ input stage 19.3µVrms @ input stage
54µV with latching 59µV with latching
Memristor tuning
range
250k covers ±700µV 280k covers ±700µV
Memristor tuning
sensivity
2µV/k
Energy (250ns) 1.44pJ 1.60pJ
Power (250ns) 5.77µW 6.39µW
isolated and will not be burned. In this step, EN MEM = 0 and EN BAR = 5V;
meanwhile, the core amplifier is off.
2. Initialisation of memristive devices. The consecutive 1µs pulses from 8Vto 12V
will be applied on the device. It will induce that resistance drops from 106 to a
more stable range of [104,105Ω].
3. Programming the device to a specific resistance. After initialisation, we can provoke
the device by a sequence of 100ns 2Vpulses to program the devices to 10kΩ. If the
resistance is read and greater than 10kΩ, we can apply reversed 2Vpulses to reduce
the resistance.
Note that the pulse will be followed by the 0.5Vtriangular wave with a duration of 50ms.
The demonstration waveform has been presented in Figure 5.3. Moreover, we expect to
programme the memristor to 1k10kΩ. The above setup and connection are visible in
Table 7.4.
7.3.2 Measurement of the Bias Current
As shown in Figure 5.2, the duration of each phase depends on the change of voltage/cur-
rent controlled by the external voltage sources. Thus, we are supposed to measure branch
currents and integrate voltages at the nodes of mida/b to adjust the ideal control signals
from the circuit simulation.
82 Chapter 7 Layout Design and Chip Testing
Table 7.4: The pin connection for programming memristive devices.
No. Pin Name Connection No. Pin Name Connection
1 GND AMPDLC gnd 12 clk rst gnd
2 clk anabar 1.8V 13 clk gnd
3 inb 1.8V 14 outb float
4 ina 1.8V 15 outa float
5 VDD AMP gnd 16 VDD DLC gnd
6 clk ana gnd 17 ENMEM 1 gnd
7 B2 2 gnd 18 ENBAR 1 5V
8 VDD MEM 5V 19 B1 1 programming pulses
9 B1 2 gnd 20 B2 1 programming pulses
10 ENBAR 2 5V 21 GND MEM gnd
11 ENMEM 2 gnd
Presented in Figure 7.2(a), terminals B1/2 allow us to measure branch currents and
voltages of mida/b when the pass transistors are turned on. In order to get the current
through the memristor, we need to connect the pads of B1/2 to the external current-
sensing device, which acquires the voltage drop across a current-sense resistor. To measure
the branch current, the core amplifier will be biased to a constant state:
clk ana = 1.8Vto allow the bias current to drain into the core amplifier.
clk anabar = 0.6Vto bias the cascode pair (M6&M7) to saturation region.
EN MEM = 5Vand ENBAR = 0 to connect memrisors to current branches of
amplifier.
clk rst = 1.8Vto turn on the reset transistors (M8&M9) to ensure the core amplifier
is in the constant status.
clk = 0 since we do not need access to DLC.
ina/b = 1Vto ensure the input signal will not incline the branch currents.
The above settings are summarised in Table 7.5. The ‘large’ signal’ of mida/b is required
to calculate the integrating time in phase (ii) in Figure 2.14. In this case, we will keep
the control signals except for the clk rst. The clk rst transitions from 1.8Vto 0 to
record the rough duration. When the voltage of mida/b increases from zero to one volt,
we record the time as t1. This helps determine the approximate phase duration, and a
more accurate time setting will be conducted in the next step. By accessing both branch
currents through the memristor and the rough integrating time, we can not determine the
phase setting but verify the function of capacitor integration and the initial error between
two branches through V=i·t
C.
Chapter 7 Layout Design and Chip Testing 83
Table 7.5: The pin connection for measuring branch current.
No. Pin Name Connection No. Pin Name Connection
1 GND AMPDLC gnd 12 clk rst 1.8V
2 clk anabar 0.6V 13 clk gnd
3 inb 1V 14 outb float
4 ina 1V 15 outa float
5 VDD AMP 1.8V 16 VDD DLC 1.8V
6 clk ana 1.8V 17 ENMEM 1 5V
7 B2 2 current sensing 18 ENBAR 1 gnd
8 VDD MEM 5V 19 B1 1 current sensing
9 B1 2 current sensing 20 B2 1 current sensing
10 ENBAR 2 gnd 21 GND MEM gnd
11 ENMEM 2 5V
7.3.3 Determining the Operation Phases
After calibrating the chip, we need to determine three operation phases: (i) reset, (ii)
integration and (iii) digitisation phase by monitoring the analogue (B2 1 &B2 2) or digital
(outa & outb) outputs.
The setup of measuring the reset phase is given in Table 7.6. In this case, we should
monitor the differential voltage of analogue outputs Vmid (from pads B2 1 &B2 2) by
using embed LPC1768. We record the duration that Vmid needs to settle down. In this
measurement, we can try sweeping the Vclk anabar and Vina/b to determine their effects on
the reset time.
Table 7.6: The pin connection for measuring the reset phase.
No. Pin Name Connection No. Pin Name Connection
1 GND AMPDLC gnd 12 clk rst 1.8V
2 clk anabar 0.6V 13 clk gnd
3 inb 1V 14 outb float
4 ina 1V+1mV 15 outa float
5 VDD AMP 1.8V 16 VDD DLC 1.8V
6 clk ana 1.8V 17 ENMEM 1 5V
7 B2 2 mbed sensing 18 ENBAR 1 gnd
8 VDD MEM 5V 19 B1 1 floating
9 B1 2 floating 20 B2 1 mbed sensing
10 ENBAR 2 gnd 21 GND MEM gnd
11 ENMEM 2 5V
Table 7.7 shows the setup of measuring the integration phase. In this measurement, we
need to switch off the bottom transistors to allow integration on the load capacitors and
recording Vmid. In this case, we need to record the below variables: (i) sweep Vclk anabar
and record Vmid versus integration time. (ii) keep Vin and sweep Vinb to measure the
84 Chapter 7 Layout Design and Chip Testing
Vmid so that we can obtain the result similar to Figure 5.4 and the common mode input
range. (iii) keep Vinb and sweep Vina to determine the differential voltage gain (similar to
Figure 5.5).
Table 7.7: The pin connection for measuring the integrating phase.
No. Pin Name Connection No. Pin Name Connection
1 GND AMPDLC gnd 12 clk rst 1.8V 0V
2 clk anabar 0.6V 13 clk gnd
3 inb 1V 14 outb float
4 ina 1V+1mV 15 outa float
5 VDD AMP 1.8V 16 VDD DLC 1.8V
6 clk ana 1.8V 17 ENMEM 1 5V
7 B2 2 mbed sensing 18 ENBAR 1 gnd
8 VDD MEM 5V 19 B1 1 floating
9 B1 2 floating 20 B2 1 mbed sensing
10 ENBAR 2 gnd 21 GND MEM gnd
11 ENMEM 2 5V
7.3.4 Voltage Gain
Since the gain depends mainly on the integration time, we need to identify the exact
proper duration for the mida/b to reach the voltage to trigger DLC and acceptable gain
with a time margin. Thus, we designed the process of assessing the gain based on the
fabricated device, taking the ideal simulations as a reference. The process accesses the
exact duration of each phase and gains as follows.
1. From capturing the voltage, we can obtain the reference time t1and gain g1. In the
test, the reset and digitisation time is based on the test result above. At first, we
set the integration phase as t1for ina = 1V+ 1mV and inb = 1V.
2. Test the differential gain gxwith the integration time of tx=t110ns and find if
the peak of gain voltage appear. If the maximum gain appears (corresponding peak
in Figure 5.5), we record the peak as gmax and time as tmax. If the gain decreases
progressively to zero, we can move to the next step.
3. Record the gain gxwith the integration time of tx=t1+10ns and find out the peak
gain gmax and tmax.
4. With the consideration of time margin, we need to set the actual integrating time as
t=tmax20ns and acceptable gain as gwith the reason in Section 4.2.3. Eventually,
we will access each phase’s gain and accurate testing time, which can be utilised in
the following testing.
Chapter 7 Layout Design and Chip Testing 85
7.3.5 Noise, CMRR and PSRR
The input-referred noise can be determined by feeding fixed input voltage to the chip for
1000 cycles and recording the digital output, where we can obtain the detection accuracy
directly. Then, we can sweep the differential input voltage to obtain a full map of detection
accuracy versus differential input voltage, followed by the Gaussian fitting to derive the
standard derivation (σ) as the desired input-referred noise. The methodology is similar
to the transient noise simulation in Section 5.2.2.
The testing of CMRR is similar to the differential mode in terms of the setting of control
signals. The control signals are pulses from the wave generator with different duty cycles
within 1µs detection period. As for the input voltage, we keep the input difference to
1mV for higher tolerance to the noise. Meanwhile, change the common voltage Vcm with
5mV per step from 0 to 1.8V. There will be five detection cycles for each Vcm, so we
can average them to obtain a more accurate gain. After acquiring a train of gain from
Vcm [0,1.8V], we can process the data using Matlab. After the analysis, it is possible
to find another common mode voltage Vcm=xthat induces more stability to the circuit.
A similar setup can be utilised to measure the PSRR.
7.3.6 Tunable Range and Sensitivity
In this case, we can test the offset tuning range and tuning sensitivity on the resistive state
of memristive devices. Considering the duty cycle in the waveform generator, we set one
detection cycle to 1µs in the tunable testing range. The testing follows the same principle
in Figure 2.15, sweeping input across a range of 2mV with 1mV to 1mV ascending and
1mV to 1mV descending at 1Vcommon voltage. The testing is within 400µs, yielding
an offset estimate resolution of 10µV . In this case, we can only assess outa/b for the
tuneable range.
As for the signal setting:
Import the DC Vcm=xto inb, while the ina will be the triangular waveform from
Vcm=x1mV to Vcm=x+ 1mV .
The control signals are the same as above.
Program the memristive devices to record the switching boundary of outa/b which
corresponds to the voltage of inb in the tuning range.
86 Chapter 7 Layout Design and Chip Testing
7.4 Precise Signal Generation Board for Chip Test
As mentioned above, to test the pre-amplifier chip, we need to generate precise analogue
signals for ina/b and record the analogue signals from mida/b and digital signals from
outa/b. The embed LPC1768 can record analogue and digital signals with 96MHz, indi-
cating the device can record the signals around every 10ns. Besides, the LPC1768 can
convert the analogue signals to 16-bit digital and operates at 3.3V, which indicates it
can record analogue signal minimum to 50µV . As for the one-bit digital signals can be
directly recorded by embed LPC1768 to present as ones or zeros. In this measurement,
we plan to input the differential voltage down to 50µV to check the circuit functionality
and linearity.
As for the precise analogue input (to ina/b), we design a signal generator that utilises
a 16-bit ADC (DAC8532) and ultra-low offset amplifier (OPA191). The schematic is
presented in Figure 7.8.
OPA191_1 OPA19 1_2
R1
40kOhm
R2
200Ohm
R3 200Ohm
R5 200Ohm
R4
40kOhm
R6
40kOhm
Vref (from DAC)
Vin (from DAC) Vtest (to mbed)
2.5V
ina (to test chip)
inb (to test chip)
Figure 7.8: The schematic of utilising two OPA191 to generate and test precise analogue
signals.
The schematic presents that the outputs of DAC (Vref &Vin) feed into OPA191 1. Vref is
directly fed into inb of the chip, while the voltage difference VDAC =VinVref is reduced
by 200 times to micron-volt level by OPA191 1, resulting Vina =Vinb + VDAC /200.
However, due to the device offset and variation, the result feeding in ina/b might be
shifted. To measure the input signal of the chip, we utilise the OPA191 2 to amplify the
VDAC by 200 times and record it in embed LPC1768 as Vtest. The principle relies on
the below derivation.
Vin = (∆VDAC +Vos,1)×R2
R1
(7.1)
where, Vin is the differential input to the chip, VDAC is the differential output from
DAC, and Vos,1is the offset from OPA191 1. Equation 7.1 can be converted as below.
VDAC = Vin ×R2
R1Vos,1(7.2)
The mbed LPC1768 record the amplified Vin as Vtest. By considering the device offset,
Chapter 7 Layout Design and Chip Testing 87
Vtest is
Vtest 2.5 = R6
R5×(∆Vin +Vos,2) (7.3)
where Vos,2is the offset of OPA191 2. There exist an error between VDAC and Vtest
2.5, and we set it as a parameter:
Verror = VDAC (∆Vtest 2.5) (7.4)
In this case, VDAC is set by the embed device and Vtest is recorded directly by the embed
device. Thus, Verror can be derived directly. Combining the Equation 7.2,7.3 and 7.4,
we obtain a function of Verror
Verror = ( R6
R5R1
R2
)×Vin + (R6
R5×Vos,2+Vos,1) (7.5)
Therefore, the differential input to the chip Vin can be reflected by the recorded signal
Verror :
Vin =Verror R6
R5×Vos,2Vos,1
R6
R5R1
R2
(7.6)
7.5 Pre-amplifier Chip Test
In this tape-out, we got the chip that stops at the Metal 4 layer from the foundry, which
allows us to assemble the memristor and memristor wiring on the top. The testing is
conducted on both chips (with and without memristors).
7.5.1 Testing on the Chip without Memristor
The full view of the chip without memristor and memristor pad is presented in Figure 7.9.
It can be seen that the memristor pad is missing, and we can only access the DLC
with floating DLC input. The available circuit which can be accessed, is presented in
Figure 7.10.
To test the DLC module of the chip, we set the pad connection as shown in Table 7.8.
However, the inputs of DLC are floating that we fail to measure the full function of the
DLC.
The measurement results are present in Figure 7.11. In this measurement, we import the
pulse train as the clk into the pre-amplifier chip, where the pulse duration is 10ms with
different pulse widths {2ms, 4ms, 6ms, 8ms}. It can be seen that the DLC can output
the digital signals based on different pulse width cycles. It indicates that the pre-amplifier
can be powered up adequately. However, the result shows a severe delay in the output,
which the probe card and jumper wires between the chip and the test board might cause.
88 Chapter 7 Layout Design and Chip Testing
Figure 7.9: Full view of the chip without the memristor electrode.
M0
M1
M3
M4
RRAM
CL
Vdd
clk_anabar (pad)
clk_rst(pad)
inb(pad)
MID
Q1
ENBAR(pad) Q2 ENMEM(pad)
Q3
ENBAR(pad) Q4 ENMEM(pad)
B2 (pad)
B1(pad)
M3
M5
M7
M9 C2
midb
DLC
outa
(pad)
outb
(pad)
clk (pad)
mida
open circuit
Figure 7.10: The schematic view presents the connection inside the chip where there
exist open circuit due to missing memristor.
In addition, Voutb swings between 0.9V and 1.1V, which the floating input might cause.
However, the reason is still unclear until we can access the nodes of mida/b.
7.5.2 Testing on the Chip with Memristor
The pre-amplifier chip with memristor and memristor electrodes (B2 2 & B2 1) on the
top is presented in Figure 7.12. However, fabricating the memristor directly on the top
shorts all Metal 4 layer, including all the power signals across the padring. It has been
tested that all the power is shorted in this chip.
Chapter 7 Layout Design and Chip Testing 89
Table 7.8: The pin connection for measuring the DLC only.
No. Pin Name Connection No. Pin Name Connection
1 GND AMPDLC gnd 12 clk rst gnd
2 clk anabar 1.8V 13 clk pulses
3 inb 1.8V 14 outb mbed monitoring
4 ina 1.8V 15 outa mbed monitoring
5 VDD AMP 1.8V 16 VDD DLC 1.8V
6 clk ana gnd 17 ENMEM 1 gnd
7 B2 2 / 18 ENBAR 1 5V
8 VDD MEM 5V 19 B1 1 /
9 B1 2 / 20 B2 1 /
10 ENBAR 2 5V 21 GND MEM gnd
11 ENMEM 2 gnd
0 5 10 15 20 25 30
t (ms)
0
0.5
1
1.5
2
Voltage (V)
0 5 10 15 20 25 30
t (ms)
0
0.5
1
1.5
2
Voltage (V)
0 5 10 15 20 25 30
t (ms)
0
0.5
1
1.5
2
Voltage (V)
0 5 10 15 20 25 30
t (ms)
0
0.5
1
1.5
2
Voltage (V)
0 5 10 15 20 25 30
t (ms)
0
0.5
1
1.5
2
Voltage (V)
0 5 10 15 20 25 30
t (ms)
0
0.5
1
1.5
2
Voltage (V)
0 5 10 15 20 25 30
t (ms)
0
0.5
1
1.5
2
Voltage (V)
0 5 10 15 20 25 30
t (ms)
0
0.5
1
1.5
2
Voltage (V)
(a) (b)
(d)
(c)
V(outa) V(outb)
V(clk)
V(outa) V(outb)
V(clk)
V(outa) V(outb)
V(clk)
V(outa) V(outb)
V(clk)
Figure 7.11: Testing Result of the DLC with floating inputs with different pulse
widths. In this test, the pulse in the duration of 10ms with four pulse widths
{2ms, 4ms, 6ms, 8ms}are fed into the chip (presented at the top trace of sub-figures).
The digital outputs are presented in the bottom trace of each sub-figure.
90 Chapter 7 Layout Design and Chip Testing
B2_2
B2_1
Figure 7.12: The pre-amplifier chip with the memristor pads and wiring on the top (in
white). But the wirings across the padring short the power.
7.5.3 Circuitry Improvements
To summarise, the chip test cannot be completed in the current stage and we can only
access the DLC with floating input. Nevertheless, we can still gain insights from this chip
testing, improve our layout, and debug the design.
1. Choose the square padring that reduces wiring between the core circuit
and pads. It can be seen that there exists a long wiring route for the in-line padring.
In order to reduce the wiring and potential parasitics, we can apply a square padring for
future tape-out.
2. Wiring for the memristors and pads. There are two options to overcome the short
circuit in the chip. The first option is to deposit an isolation layer to cover the padring
before having direct wiring as Figure 7.12 shown. The second option is to develop a new
wiring route that avoids the Metal 4 layer to connect the memristors and pads, presented
in Figure 7.13. Future testing will be conducted on this chip.
Figure 7.13: The full layout with new wiring route for memristors and electrodes that
prevents shorting circuit.
3. Reduce jumper wires in chip testing. A severe delay on the DLC might be caused
by the long wiring between the chip and the test board. To solve this, we can assemble
the pre-amplifier chip in the package PLCC68. Then, the PLCC68 can be assembled on
the test board and we can directly conduct the testing on the packed chip. The wire
bonding in PLCC68 presents in Figure 7.14.
Chapter 7 Layout Design and Chip Testing 91
123456789 68 67 66 65 64 63 62 61
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
353433323130292827 36 37 38 39 40 41 42 43
60
59
58
57
56
55
54
53
52
51
50
49
48
47
46
45
44
Figure 7.14: The full layout with new wiring route that prevents shorting circuit.
7.6 Summary
This chapter covers (i) the layout design and post-layout simulation, (ii) preparation for
the chip test and (iii) test results. However, we cannot test the full functionality of the
preamplifier at this point due to the chip’s open circuit (missing memristors). However,
we can still gain insights from testing on DLC only and develop strategies for improving
the layout design and test strategy going forward. The improvements of the circuit should
be based on the testing results afterwards.
Chapter 8
A Memristor-Based ∆ΣADC
In this chapter, we introduce the ∆ΣADC structure based on the initial design of the
front-end to reduce the input-referred noise through over-sampling and noise shaping.
The proposed ∆ΣADC utilises the initial design as the Gm-C integrator and induces a
current-steering DAC (IDAC) as an extra module. The architecture of the ∆ΣADC is
presented in Figure 8.1. The front-end consists of two loops: (i) a DC cancellation loop
with memristive devices and (ii) a delta-sigma modulator loop filter to shape the noise to
high frequencies.
Vip
Vin
DLC
+
-
Voutp
Voutn
Vop
D/A
Gm ArC
One
Von
Figure 8.1: The block diagram of proposed ∆ΣADC.
The proposed circuit utilises the ArC One as the external device to program the memristor.
ArC One is the platform to validate and calibrate the memristor [98]. The operation of
the ∆ΣADC can be divided into two stages. Stage one is to block the delta-sigma loop,
followed by the ArCOne instrument programs the memristor and cancels the offset based
on the output of DLC. The ArCOne stops programming when the status of Voutp and
Voutn switch, which indicates the DC offset is compensated. The second stage is to block
the Offset modulation loop and turn on the delta-sigma loop to detect the neural spike
with low noise.
92
Chapter 8 A Memristor-Based ∆ΣADC 93
8.1 Architecture of ∆ΣADC
In this thesis, we proposed Gm-C based one-bit continuous-time delta-sigma (CT-∆Σ)
ADC modulator. The CT techniques feature an inherent anti-aliasing filter, which re-
duces system components and energy consumption. However, the first-stage integrator
occupies a large percentage of overall power. As for the integrator, the active-RC inte-
grator produces high linearity at the cost of more power required to run the feedback
network with enough loop gain and phase margin. On the other hand, the Gm-C integra-
tor utilises less power than the active-RC integrator while driving an open capacitive load
loop. The most challenging aspect of developing a Gm-C-based modulator is the limited
linear input range caused by the nonlinear Gm amplifier, which results in a low SNDR.
Nevertheless, this system is applied for the neural signal that input changes are in the
millivolt range. It enables the development of a single-stage OTA without compromising
power to improve linearity.
M0
M1 M2
M3 M4
M5 M6
M7 M8
M9
M10 M11
M12 M13
I1
I2
Vdd Vdd
R1 R2
C1
Vip Vin
Vb1
Vb2
Voutp Voutn
Vb3
VopVon
Vop Von
(a) (b)
Figure 8.2: The schematic of OTA of ∆ΣADC. (a) The OTA is the input and the
integrator interface in the ∆Σ modulator. (b) The IDAC shown on the right creates a
current feedback loop.
The core of ∆Σ modulator is built with a Gm-C integrator, current-steering DAC (IDAC),
and a DLC (shown in Figure 8.1 and Figure 8.2). The sizes of OTA are presented in
Table 8.1 while the sizes of DLC are the same as the previous design in Chapter 5. In this
case, we utilise the oversampling ratio (OSR) of 512 to achieve low quantisation error with
the signal bandwidth of 10 kHz. The OTA operates as the input stage of the channel and
94 Chapter 8 A Memristor-Based ∆ΣADC
the integrator interfaces. The OTA consists of a current source (M0), input differential
pair (M1&M2), the active load (M7&M8), and the cascode pairs (M3-M6). It operates
under the bias current of 3 µA. Its output is coupled to the load capacitor (C1), IDAC
and DLC. A high output impedance is essential to preserve the noise transfer function
of the modulator. The output is equipped with cascode pairs (M3-M7) to increase the
output impedance.
Table 8.1: Sizes of devices of OTA in the proposed ∆Σ ADC.
Devices W/L (µm/µm) Devices W/L (µm/µm)
M1, M2 8.5/6 M10, M11 1/5
M3, M4 2/2 M12, M13 1/5
M5, M6 2/5 C1 200 f F
M7, M8 1/18 I1 3µA
M0, M9 3/1 I2 240nA
In this design, we utilise the feedback path constructed in the current domain to eliminate
the need for an additional pair of input transistors. In order to minimise the area even
further, asymmetric current feedback is achieved by utilising the current source solely. The
input of the IDAC is the bitstream output by the quantiser (Voutp&Voutn). A common logic
block generates the control signals for these two blocks from the quantisation bitstream.
The IDAC feeds the regulated current of 240 nA to the output of the OTA (Vop&Von) in
a full-scale mode of 10 mV .
Similar to the above front-end designs, the proposed ∆ΣADC induces a pair of memristive
devices (R1&R2) to compensate for the DC offset. The programming module is the same
as above. In this case, the memristors are put above the input transistors, operating as
source degeneration. Meanwhile, it can compensate the same DC offset (50 mV ) with a
lower resistive state. The offset compensation is summarised in Table 8.2.
Table 8.2: The compensated DC offset versus the resistive state of memristors under
the condition of R1 = 1 kΩ.
R2 (kΩ) Offset Compensation (mV )
5 8.24
10 18.70
15 28.87
20 39.00
25 49.00
8.2 Performance ∆ΣADC
The operation of the ∆ΣADC is given in Figure 8.3 under the sine 10 mVpp input. The
feedback modulation on the nodes of Vop and Von is given by the IDAC. After collecting the
output data, we conduct the Fast Fourier Transform (FFT) to obtain the output spectrum
Chapter 8 A Memristor-Based ∆ΣADC 95
0 50 100 150
895
900
905
Voltage (mV)
0 50 100 150
0.8
1
1.2
Voltage (V)
0 50 100 150
0
1
2
Voltage (V)
0 50 100 150
Time ( s)
0
1
2
Voltage (V)
V(Vip)V(Vin)
V(Vop)V(Von)
V(Voutp)
V(Voutn)
(a)
(b)
(c)
(d)
Figure 8.3: The timing diagram of ∆ΣADC. (a) the input of the system. (b) the output
of the OTA. The one-bit IDAC provides the modulation. (c) and (d) are the outputs of
the DLC, respectively.
in Figure 8.4(a) with the noise bandwidth of [0.1Hz,100MHz]. The spectrum presents that
the noise is shaped to a higher frequency range so that the front-end can keep the low
noise level in the signal band. It can achieve 60.22 dBc Spurious-Free Dynamic Range
(SFDR) and 53dB Signal-to-Noise And Distortion Ratio (SINAD), which indicates 8.51
Effective Number Of Bits (ENOB). In addition, Figure 8.4(b) presents the SNR versus
the input signal in dB, which indicates the 55 dB Dynamic Range (DR) of the system
(AC input range). As for the DC input range, it is measured at the input differential pair
that the input voltage keeps the input transistor in saturation/sub-saturation region. The
accepted DC input range in this design is 200 mV from 0.85V to 1.05V. The performance
is summarised in Table 8.3.
The transconductance amplifier acts as the input stage for the whole system, making its
noise performance crucial. The input-referred noise of the OTA is 9.56 µVrms, mainly
given by the input differential and the load pairs of the OTA, which occupies the most
area of ∆ΣADC as well. The transistor dimensions and noise contribution are given in
Table 8.4.
96 Chapter 8 A Memristor-Based ∆ΣADC
102104106
Frequency (Hz)
-125
-100
-75
-50
-25
Output PSD (dB)
-60 -40 -20 0
Input Level (dB)
0
10
20
30
40
50
60
SNR (dB)
0
Figure 8.4: Performance of ∆ΣADC. (a) the output spectrum for the input sine wave
of 10 mVpp under the noise simulation. (b) SNR versus input level where indicates 55dB
DR.
Table 8.3: Performance of the proposed ∆ΣADC under the condition of transient noise.
Parameter Value
Fs10.24 MHz
Rout @Fs3.10 M
SINAD 53.00 dB
ENOB 8.51 bits
SFDR 60.22 dBc
Input-referred noise of OTA 9.56 µVrms
DC Input Range 0.85-1.05 V
AC Input Range 10 mV
Power 5.50 µW
Table 8.4: The critical transistor dimensions and noise contribution of the OTA.
Transistor W/L (µm/µm) Noise Contribution
M7&M8 18/1 50.2% 4.86 µVrms
M1&M2 8.5/6 48.78% 4.66 µVrms
8.3 Simulation with Spike Train
In this section, we investigate the performance of ∆Σ ADC on a spike train with 15.77mV
DC offset by conducting a transient simulation with a spike train. The timing diagram
presents in Figure 8.5, where (a) shows the spike train in inp and inn is set to 0.9V. The
OTA is operated as a conventional amplifier in this design without switching status. As
a result, the input signal is free of kickback noise (compared with the inputs in both AC-
coupled and DC-coupled front-end). In this simulation, we set R1 = 1k and R2 = 8.5k
Chapter 8 A Memristor-Based ∆ΣADC 97
to compensate 15.77mV DC offset. In addition, we get a spectrum from outn presented in
Figure 8.5. In this simulation, the input frequency is 10.07kH z and its amplitude is 500µV
at its maximum. ENOB = 1.34 indicates the resolution is 198µV , significantly lower than
the simulation of pure sine waves (ENOB = 5.33). The DC offset (15.77mV ) needs to
be compensated by high precision memristors of 0.01mV . Nevertheless, the memristors
in this circuit are placed at the source of input transistors to achieve a wider tuneable
offset range, albeit at the expense of precision. Due to the DC input, the SINAD and
ENOB may experience attenuation. Thus, improving the ∆Σ ADC in terms of eliminating
common mode input effect and practical detection remains a future work.
102103104105106107
Frequency (Hz)
-80
-70
-60
-50
-40
-30
-20
Output PSD (dB)
Freq = 10.07kHz
SINAD = 9.83dB
ENOB = 1.34bits
SFDR = 14.14dB
0 10 20 30 40 50 60 70 80 90 100
0.9155
0.916
0.9165
Voltage (V)
0 10 20 30 40 50 60 70 80 90 100
0.9
1
1.1
1.2
1.3
Voltage (V)
0 10 20 30 40 50 60 70 80 90 100
0
1
2
Voltage (V)
0 10 20 30 40 50 60 70 80 90 100
Time ( s)
0
1
2
Voltage (V)
(a)
(b)
(c)
(d)
V(inp)
V(op) V(on)
V(outp)
V(outn)
Figure 8.5: Left: the timing diagram of ∆ΣADC with (a) the spike train input with
15.77mV DC offset. (b) the output of the OTA. (c) and (d) are the outputs of the DLC,
respectively. Right: output spectrum for spike train with the maximum amplitude of
500µV and 15.77mV DC offset.
8.4 Summary
This chapter proposes the DC-coupled continuous-time ∆ΣADC based on our previous
designs. Even though the design of Gm-C based CT-∆ΣADC meets the challenge of
relative low linearity and SNDR, this limitation can be lifted since it is applied for low
input range bio-signal (10mV ). Meanwhile, memristive devices can compensate for the
large DC offset at the input. Under transient noise in the range of [1Hz,100M Hz], it
operates as an eight-bit ADC at the front-end. The OTA occupied the most significant
proportion of the total power consumption, contributing to a low input-referred noise
level achieved by the modulator’s loop filter. Compared with the front-end in Chapter 4
- Chapter 6, the OTA operates in a static mode that consumes more power than the
dynamic mode, which uses the start-stop scheme to reduce the power down to microvolt
levels. In summary, the memristor-based amplifier can be applied to the dynamic pre-
amplifier and the static one in ∆Σ ADC. This contributes to rejecting the DC offset with
low power consumption and area efficiency.
Chapter 9
Conclusion and Future Work
9.1 Conclusion
This dissertation performs a theoretical analysis of the core functionality of memristive in-
tegrating amplifiers. It uses industrial CAD-level simulations to provide a specific example
of an integrating amplifier design targeting electrophysiological applications. Throughout
our analysis, the performance enhancement over traditional, continuous-mode amplifiers
could be most intuitively understood as a gain-boosting effect arising from the integration
process. Moreover, we explained how standard amplifier performance metrics, such as gain
and input common mode range, and new metrics, such as offset voltage tuneability range,
can be described by governing equations for use by designers. Finally, we implemented
an exemplar design in commercially available 180nm CMOS and demonstrated typical
values for all studied performance parameters that can be expected from a 0.18µm node
technology.
The memristive device was utilised as a tunable resistor along the differential current
path to compensate for the offset voltage. We also studied the impact of memristor IV
non-linearity on the system, especially for the integration process. We checked whether
transitioning from ideal resistor-based analysis to realistic memristor-based presents a
fundamental conceptual challenge towards the viability of low-power, ultra-low offset
memristive amplifiers. The result was that it still needs to remove the last fundamen-
tal/conceptual roadblock we have identified towards scale-up and, ultimately, commercial
exploitation; the unique challenges are of an engineering nature.
This work is a stepping stone towards de-risking and documenting the memristor-based
integrating amplifier. With a nano-scaled memristor implanted, the circuit does not need
a large module to eliminate the offset voltage after fabrication. However, the memristive
device still has some limitations. The in-house memristor is connected to a switch unit
which needs 8Vto 12Vto switch the resistance state. In this case, the switch unit induces
noise risk to this sensitive system and increases power consumption. The trade-off induced
98
Chapter 9 Conclusion and Future Work 99
by the integration process, combined with the offset trimming enabled by the memristor,
has the potential to add a robust circuit topology to the arsenal of the analogue designer.
After defining and measuring the key performance metric for integrating amplifier, we
applied the memristor-based pre-amplifier into three types of application circuits: (i)
open-loop network with input capacitors and MOS-bipolar pseudo-resistor as an AC-
coupled solution to reject DC offset, (ii) directly operates in the DC-coupled solution
for DC offset up to 50mV , and (iii) memristor-based ∆Σ ADC. Through measurement,
the AC-coupled and DC-coupled memristor-based pre-amplifier with a ‘start-stop’ scheme
can achieve above 95% detection accuracy while maintaining the power consumption at
hundreds of nanowatt per channel. Moreover, the ∆Σ ADC with static pre-amplifier
operates at 10.24M H z, consuming 5.5µW .
In the tape-out of single channel pre-amplifier, we used in-house memristor electrodes/-
pads that directly connect to the memristor. It is because electroforming the device
requires 8 12Vthat exceeds the maximum tolerance voltage of the chosen commercial
pad. However, the latest version of the hybrid CMOS/memristor chip encounters the
situation that the memristor wiring shorts the power path that cannot be tested. Under
this situation, the chip test can only be conducted on the chip with missing memristors
and memristive pads, where an open circuit exists inside the core amplifier. We can only
access the DLC module with floating inputs. The test results indicate that the chip can
be powered up, but the DLC operates with long delays, which might be due to the long
wiring (i) inside the chip, and (ii) jumper wires between the probe card to the test board.
We can still gain some insights from this chip test regarding the test strategy and layout
design concluded in Chapter 7. Furthermore, the improvement of the circuit design based
on chip tests remains future work.
9.2 Recommendation for Future Work
The research work in this thesis brings new application prospects for the memristive
device, utilising it as the trimming component to reject the DC offset at the input in
the application of neural signal detection. The presented work can be developed in the
following ways.
1. Development of the memristive Verilog-A model. Currently, the memristive
Verilog-A model in Cadence can only trace the RS in transient simulation. Moreover, it
operates as the ideal static resistor in DC/AC simulation (without noise), which restricts
us from exploring the frequency impact on the memristor. Therefore, developing the AC
and noise model for the memristor is recommended, which contributes to higher accuracy
of the performance of the memristor in CMOS circuits. In addition, the neural front-end
should be operated with ultra-low power consumption, especially for the electroforming
and the programming of the memristive devices. In this project, we utilised the device
100 Chapter 9 Conclusion and Future Work
that requires 8 12Vfor electroforming/activation. It induces the risk of destroying the
circuit in 0.18µm even though we utilised a memristor electrode and separated it from
other signal pads. For future neural front-end development, developing the Verilog-A
model for low-voltage activation is recommended, which contributes to power efficiency.
2. Develop the automated programming interface of the memristor. Currently,
we utilise the memristor for trimming the offset voltage by monitoring the digital out-
put and manually programming using an external device (ArCOne Instrument). It is
recommended to develop an automated programming interface that incorporates a detec-
tion/comparison circuit at the output of the OTA to control the programming module
until the DC offset is completely covered. Note that this can be realised when the mem-
ristor can be electroformed and programmed with a low voltage. In this case, the circuit
can be calibrated at the first stage, followed by the actual operation.
3. Full system development. Current research focuses on the design of a single-
channel pre-amplifier. The design of multiple channels and implementation of the full
system remain future work. The focus should be placed on parasitic analysis in the multi-
channel design. Moreover, a power module is required to generate stable bias current
to different pre-amplifier channels. In addition, the full system requires a controlling
module that generates clocking signals to control the operation of the pre-amplifier with
the consideration of clock skew and jitter.
Bibliography
[1] J. Kao, S. Stavisky, D. Sussillo, P. Nuyujukian, and K. Shenoy, “Information systems
opportunities in brain–machine interface decoders,” Proceedings of the IEEE, vol.
102, pp. 666–682, 2014.
[2] S. B. Lee, H.-M. Lee, M. Kiani, U.-M. Jow, and M. Ghovanloo, “An inductively pow-
ered scalable 32-channel wireless neural recording system-on-a-chip for neuroscience
applications,” IEEE transactions on biomedical circuits and systems, vol. 4, no. 6,
pp. 360–371, 2010.
[3] S. Nag and N. V. Thakor, “Implantable neurotechnologies: electrical stimulation
and applications,” Medical & biological engineering & computing, vol. 54, no. 1, pp.
63–76, 2016.
[4] D. Loi, C. Carboni, G. Angius, G. N. Angotzi, M. Barbaro, L. Raffo, S. Raspopovic,
and X. Navarro, “Peripheral neural activity recording and stimulation system,” IEEE
transactions on biomedical circuits and systems, vol. 5, no. 4, pp. 368–379, 2011.
[5] D. Tsai, D. Sawyer, A. Bradd, R. Yuste, and K. L. Shepard, “A very large-scale mi-
croelectrode array for cellular-resolution electrophysiology,” Nature communications,
vol. 8, no. 1, pp. 1–11, 2017.
[6] X. Yuan, M. Schr¨oter, M. E. J. Obien, M. Fiscella, W. Gong, T. Kikuchi, A. Odawara,
S. Noji, I. Suzuki, J. Takahashi et al., “Versatile live-cell activity analysis platform
for characterization of neuronal dynamics at single-cell and network level,” Nature
communications, vol. 11, no. 1, pp. 1–14, 2020.
[7] S. Wang, S. K. Garakoui, H. Chun, D. G. Salinas, W. Sijbers, J. Putzeys, E. Martens,
J. Craninckx, N. Van Helleputte, and C. M. Lopez, “A compact quad-shank cmos
neural probe with 5,120 addressable recording sites and 384 fully differential parallel
channels,” IEEE transactions on biomedical circuits and systems, vol. 13, no. 6, pp.
1625–1634, 2019.
[8] R. Fi´ath, B. C. Raducanu, S. Musa, A. Andrei, C. M. Lopez, M. Welkenhuysen,
P. Ruther, A. Aarts, and I. Ulbert, “Fine-scale mapping of cortical laminar activity
during sleep slow oscillations using high-density linear silicon probes,” Journal of
neuroscience methods, vol. 316, pp. 58–70, 2019.
101
102 BIBLIOGRAPHY
[9] G. N. Angotzi, F. Boi, A. Lecomte, E. Miele, M. Malerba, S. Zucca, A. Casile, and
L. Berdondini, “Sinaps: An implantable active pixel sensor cmos-probe for simul-
taneous large-scale neural recordings,” Biosensors and Bioelectronics, vol. 126, pp.
355–364, 2019.
[10] E. Musk and Neuralink, “An integrated brain-machine interface platform with thou-
sands of channels (submitted),” 2019.
[11] X. Zou, L. Liu, J. H. Cheong, L. Yao, P. Li, M.-Y. Cheng, W. L. Goh, R. Rajkumar,
G. S. Dawe, K.-W. Cheng et al., “A 100-channel 1-mw implantable neural recording
ic,” IEEE Transactions on Circuits and Systems I: Regular Papers, vol. 60, no. 10,
pp. 2584–2596, 2013.
[12] J. Ruiz-Amaya, A. Rodriguez-Perez, and M. Delgado-Restituto, “A low noise ampli-
fier for neural spike recording interfaces,” Sensors, vol. 15, no. 10, pp. 25 313–25 335,
2015.
[13] M. Yin and M. Ghovanloo, “A low-noise preamplifier with adjustable gain and band-
width for biopotential recording applications,” in 2007 IEEE International Sympo-
sium on Circuits and Systems. IEEE, 2007, pp. 321–324.
[14] H. Gao, R. M. Walker, P. Nuyujukian, K. A. Makinwa, K. V. Shenoy, B. Murmann,
and T. H. Meng, “Hermese: A 96-channel full data rate direct neural interface in
0.13µm cmos,” IEEE Journal of Solid-State Circuits, vol. 47, no. 4, pp. 1043–1055,
2012.
[15] M. S. Chae, Z. Yang, M. R. Yuce, L. Hoang, and W. Liu, “A 128-channel 6 mw
wireless neural recording ic with spike feature extraction and uwb transmitter,” IEEE
transactions on neural systems and rehabilitation engineering, vol. 17, no. 4, pp. 312–
321, 2009.
[16] R. R. Harrison, P. T. Watkins, R. J. Kier, R. O. Lovejoy, D. J. Black, B. Greger,
and F. Solzbacher, “A low-power integrated circuit for a wireless 100-electrode neural
recording system,” IEEE Journal of Solid-State Circuits, vol. 42, no. 1, pp. 123–133,
2006.
[17] A. D. Rush and P. R. Troyk, “A power and data link for a wireless-implanted neural
recording system,” IEEE Transactions on Biomedical Engineering, vol. 59, no. 11,
pp. 3255–3262, 2012.
[18] R. Sharpeshkar, “Ultra low power bioelectronics: Fundamentals, biomedical applica-
tions, and bio-inspired system,” 2010.
[19] I. Karageorgos, K. Sriram, J. Vesel`y, M. Wu, M. Powell, D. Borton, R. Manohar,
and A. Bhattacharjee, “Hardware-software co-design for brain-computer interfaces,”
in 2020 ACM/IEEE 47th Annual International Symposium on Computer Architecture
(ISCA). IEEE, 2020, pp. 391–404.
BIBLIOGRAPHY 103
[20] N. Even-Chen, D. G. Muratore, S. D. Stavisky, L. R. Hochberg, J. M. Henderson,
B. Murmann, and K. V. Shenoy, “Power-saving design opportunities for wireless in-
tracortical brain–computer interfaces,” Nature biomedical engineering, vol. 4, no. 10,
pp. 984–996, 2020.
[21] J. T. Robinson, E. Pohlmeyer, M. C. Gather, C. Kemere, J. E. Kitching, G. G.
Malliaras, A. Marblestone, K. L. Shepard, T. Stieglitz, and C. Xie, “Developing
next-generation brain sensing technologies—a review,” IEEE sensors journal, vol. 19,
no. 22, pp. 10 163–10 175, 2019.
[22] I. Gupta, A. Serb, A. Khiat, R. Zeitler, S. Vassanelli, and T. Prodromakis, “Real-time
encoding and compression of neuronal spikes by metal-oxide memristors,” Nature
communications, vol. 7, no. 1, pp. 1–9, 2016.
[23] L. Fassio, L. Lin, R. De Rose, M. Lanuzza, F. Crupi, and M. Alioto, “Trimming-
less voltage reference for highly uncertain harvesting down to 0.25 v, 5.4 pw,” IEEE
Journal of Solid-State Circuits, vol. 56, no. 10, pp. 3134–3144, 2021.
[24] S. Lv, J. Liu, and Z. Geng, “Application of memristors in hardware security: A
current state-of-the-art technology,” Advanced Intelligent Systems, vol. 3, no. 1, p.
2000127, 2021.
[25] Q. Xia, W. Robinett, M. W. Cumbie, N. Banerjee, T. J. Cardinali, J. J. Yang, W. Wu,
X. Li, W. M. Tong, D. B. Strukov et al., “Memristor- cmos hybrid integrated circuits
for reconfigurable logic,” Nano letters, vol. 9, no. 10, pp. 3640–3645, 2009.
[26] F. Cai, J. M. Correll, S. H. Lee, Y. Lim, V. Bothra, Z. Zhang, M. P. Flynn, and
W. D. Lu, “A fully integrated reprogrammable memristor–cmos system for efficient
multiply–accumulate operations,” Nature Electronics, vol. 2, no. 7, pp. 290–299, 2019.
[27] S. Stathopoulos, A. Khiat, M. Trapatseli, S. Cortese, A. Serb, I. Valov, and T. Pro-
dromakis, “Multibit memory operation of metal-oxide bi-layer memristors,” Scientific
reports, vol. 7, no. 1, pp. 1–7, 2017.
[28] P. Mazumder, S.-M. Kang, and R. Waser, “Memristors: devices, models, and appli-
cations,” Proceedings of the IEEE, vol. 100, no. 6, pp. 1911–1919, 2012.
[29] K.-H. Kim, S. Gaba, D. Wheeler, J. M. Cruz-Albrecht, T. Hussain, N. Srinivasa, and
W. Lu, “A functional hybrid memristor crossbar-array/cmos system for data storage
and neuromorphic applications,” Nano letters, vol. 12, no. 1, pp. 389–395, 2012.
[30] A. Serb and T. Prodromakis, “High-sensitivity memristor-based threshold detection,”
in 2018 IEEE International Symposium on Circuits and Systems (ISCAS). IEEE,
2018, pp. 1–4.
[31] J. Zha, H. Huang, T. Huang, J. Cao, A. Alsaedi, and F. E. Alsaadi, “A general mem-
ristor model and its applications in programmable analog circuits,” Neurocomputing,
vol. 267, pp. 134–140, 2017.
104 BIBLIOGRAPHY
[32] Q. Xia, “Memristive nanodevices: Cmos compatibility and novel applications,” in
2016 18th Mediterranean Electrotechnical Conference (MELECON). IEEE, 2016,
pp. 1–4.
[33] I. Bhati, M.-T. Chang, Z. Chishti, S.-L. Lu, and B. Jacob, “Dram refresh mechanisms,
penalties, and trade-offs,” IEEE Transactions on Computers, vol. 65, no. 1, pp. 108–
121, 2015.
[34] K. R. GENDEPUJARI, “Cantilever based nonvolatile memories: A survey,” Ph.D.
dissertation, SARDAR VALLABHBHAI NATIONAL INSTITUTE OF TECHNOL-
OGY, SURAT, 2020.
[35] S. Maheshwari, S. Stathopoulos, J. Wang, A. Serb, Y. Pan, A. Mifsud, L. B.
Leene, J. Shen, C. Papavassiliou, T. G. Constandinou et al., “Design flow for hybrid
cmos/memristor systems—part i: Modeling and verification steps,” IEEE Transac-
tions on Circuits and Systems I: Regular Papers, vol. 68, no. 12, pp. 4862–4875,
2021.
[36] I. Messaris, S. Nikolaidis, A. Serb, S. Stathopoulos, I. Gupta, A. Khiat, and T. Pro-
dromakis, “A tio2 reram parameter extraction method,” in 2017 IEEE International
Symposium on Circuits and Systems (ISCAS). IEEE, 2017, pp. 1–4.
[37] I. Messaris, A. Serb, S. Stathopoulos, A. Khiat, S. Nikolaidis, and T. Prodromakis, “A
data-driven verilog-a reram model,” IEEE Transactions on Computer-Aided Design
of Integrated Circuits and Systems, vol. 37, no. 12, pp. 3151–3162, 2018.
[38] J. Wang, A. Serb, C. Papavassiliou, S. Maheshwari, and T. Prodromakis, “Analysing
and measuring the performance ofmemristive integrating amplifiers,” arXiv preprint
arXiv:2009.09909, 2020.
[39] J. Wang, A. Serb, S. Wang, and T. Prodromakis, “Hybrid cmos/memristor front-
end for multiunit activity processing,” in 2022 IEEE International Symposium on
Circuits and Systems (ISCAS). IEEE, 2022, pp. 1–5.
[40] J. Wang, A. Serb, C. Papavassiliou, and T. Prodromakis, “Accounting for memristor
iv non-linearity in low power memristive amplifiers,” in 2021 IEEE International
Symposium on Circuits and Systems (ISCAS). IEEE, 2021, pp. 1–5.
[41] J. Wang, A. Serb, S. Wang, and T. Prodromakis, “Offset rejection in a dc-coupled
hybrid cmos/memristor neural front-end,” in 2022 IEEE International Symposium
on Circuits and Systems (ISCAS). IEEE, 2022, pp. 1–5.
[42] B. P. Bean, “The action potential in mammalian central neurons,” Nature Reviews
Neuroscience, vol. 8, no. 6, pp. 451–465, 2007.
BIBLIOGRAPHY 105
[43] M. Dipalo, H. Amin, L. Lovato, F. Moia, V. Caprettini, G. C. Messina, F. Tantussi,
L. Berdondini, and F. De Angelis, “Intracellular and extracellular recording of spon-
taneous action potentials in mammalian neurons and cardiac cells with 3d plasmonic
nanoelectrodes,” Nano letters, vol. 17, no. 6, pp. 3932–3939, 2017.
[44] E. R. Kandel, J. H. Schwartz, T. M. Jessell, S. Siegelbaum, A. J. Hudspeth, S. Mack
et al.,Principles of neural science. McGraw-hill New York, 2000, vol. 4.
[45] B. Gosselin, “Recent advances in neural recording microsystems,” Sensors, vol. 11,
no. 5, pp. 4572–4597, 2011.
[46] G. Dornhege, J. d. R. Mill´an, T. Hinterberger, D. J. McFarland, K.-r. Muller et al.,
Toward brain-computer interfacing. Citeseer, 2007, vol. 63.
[47] A. Molleman, Patch clamping: an introductory guide to patch clamp electrophysiol-
ogy. John Wiley & Sons, 2003.
[48] J. G. Webster, Medical instrumentation: application and design. John Wiley &
Sons, 2009.
[49] P. G. Patil and D. A. Turner, “The development of brain-machine interface neuro-
prosthetic devices,” Neurotherapeutics, vol. 5, no. 1, pp. 137–146, 2008.
[50] Y.-S. Choi, M. A. Koenig, X. Jia, and N. V. Thakor, “Quantifying time-varying mul-
tiunit neural activity using entropy-based measures,” IEEE transactions on biomed-
ical engineering, vol. 57, no. 11, pp. 2771–2777, 2010.
[51] E. Stark and M. Abeles, “Predicting movement from multiunit activity,” Journal of
Neuroscience, vol. 27, no. 31, pp. 8387–8394, 2007.
[52] I. Stolerman, Encyclopedia of psychopharmacology. Springer Science & Business
Media, 2010.
[53] G. Hong and C. M. Lieber, “Novel electrode technologies for neural recordings,”
Nature Reviews Neuroscience, vol. 20, no. 6, pp. 330–345, 2019.
[54] T. Torfs, A. A. Aarts, M. A. Erismis, J. Aslam, R. F. Yazicioglu, K. Seidl, S. Herwik,
I. Ulbert, B. Dombovari, R. Fiath et al., “Two-dimensional multi-channel neural
probes with electronic depth control,” IEEE Transactions on Biomedical Circuits
and Systems, vol. 5, no. 5, pp. 403–412, 2011.
[55] A. C. Paulk, Y. Kfir, A. R. Khanna, M. L. Mustroph, E. M. Trautmann, D. J. Soper,
S. D. Stavisky, M. Welkenhuysen, B. Dutta, K. V. Shenoy et al., “Large-scale neural
recordings with single neuron resolution using neuropixels probes in human cortex,”
Nature Neuroscience, vol. 25, no. 2, pp. 252–263, 2022.
106 BIBLIOGRAPHY
[56] N. A. Steinmetz, C. Aydin, A. Lebedeva, M. Okun, M. Pachitariu, M. Bauza,
M. Beau, J. Bhagat, C. ohm, M. Broux et al., “Neuropixels 2.0: A miniatur-
ized high-density probe for stable, long-term brain recordings,” Science, vol. 372, no.
6539, p. eabf4588, 2021.
[57] S. Wurth, M. Capogrosso, S. Raspopovic, J. Gandar, G. Federici, N. Kinany,
A. Cutrone, A. Piersigilli, N. Pavlova, R. Guiet et al., “Long-term usability and
bio-integration of polyimide-based intra-neural stimulating electrodes,” Biomateri-
als, vol. 122, pp. 114–129, 2017.
[58] G. Hong, X. Yang, T. Zhou, and C. M. Lieber, “Mesh electronics: a new paradigm for
tissue-like brain probes,” Current opinion in neurobiology, vol. 50, pp. 33–41, 2018.
[59] X. Yang, T. Zhou, T. J. Zwang, G. Hong, Y. Zhao, R. D. Viveros, T.-M. Fu, T. Gao,
and C. M. Lieber, “Bioinspired neuron-like electronics,” Nature materials, vol. 18,
no. 5, pp. 510–517, 2019.
[60] T.-M. Fu, G. Hong, R. D. Viveros, T. Zhou, and C. M. Lieber, “Highly scalable
multichannel mesh electronics for stable chronic brain electrophysiology,” Proceedings
of the National Academy of Sciences, vol. 114, no. 47, pp. E10 046–E10 055, 2017.
[61] J. K. Krauss, N. Lipsman, T. Aziz, A. Boutet, P. Brown, J. W. Chang, B. Davidson,
W. M. Grill, M. I. Hariz, A. Horn et al., “Technology of deep brain stimulation:
current status and future directions,” Nature Reviews Neurology, vol. 17, no. 2, pp.
75–87, 2021.
[62] H. Shin, Y. Son, U. Chae, J. Kim, N. Choi, H. J. Lee, J. Woo, Y. Cho, S. H.
Yang, C. J. Lee et al., “Multifunctional multi-shank neural probe for investigating
and modulating long-range neural circuits in vivo,” Nature communications, vol. 10,
no. 1, pp. 1–11, 2019.
[63] M. Zhang, Z. Tang, X. Liu, and J. Van der Spiegel, “Electronic neural interfaces,”
Nature Electronics, vol. 3, no. 4, pp. 191–200, 2020.
[64] J. N. Aziz, K. Abdelhalim, R. Shulyzki, R. Genov, B. L. Bardakjian, M. Derchansky,
D. Serletis, and P. L. Carlen, “256-channel neural recording and delta compression
microsystem with 3d electrodes,” IEEE Journal of Solid-State Circuits, vol. 44, no. 3,
pp. 995–1005, 2009.
[65] D. Y. Barsakcioglu, Y. Liu, P. Bhunjun, J. Navajas, A. Eftekhar, A. Jackson, R. Q.
Quiroga, and T. G. Constandinou, “An analogue front-end model for developing
neural spike sorting systems,” IEEE transactions on biomedical circuits and systems,
vol. 8, no. 2, pp. 216–227, 2014.
[66] H. G. Rey, C. Pedreira, and R. Q. Quiroga, “Past, present and future of spike sorting
techniques,” Brain research bulletin, vol. 119, pp. 106–117, 2015.
BIBLIOGRAPHY 107
[67] C. D. Ferris, Introduction to bioinstrumentation. Humana Press, 1978.
[68] D. De Dorigo, C. Moranz, H. Graf, M. Marx, D. Wendler, B. Shui, A. S. Herbawi,
M. Kuhl, P. Ruther, O. Paul et al., “Fully immersible subcortical neural probes
with modular architecture and a delta-sigma adc integrated under each electrode for
parallel readout of 144 recording sites,” IEEE Journal of Solid-State Circuits, vol. 53,
no. 11, pp. 3111–3125, 2018.
[69] D.-Y. Yoon, S. Pinto, S. Chung, P. Merolla, T.-W. Koh, and D. Seo, “A 1024-channel
simultaneous recording neural soc with stimulation and real-time spike detection,”
in 2021 Symposium on VLSI Circuits. IEEE, 2021, pp. 1–2.
[70] X. T. Pham, N. T. Nguyen, V. T. Nguyen, and J.-W. Lee, “A 0.6-µw chopper
amplifier using a noise-efficient dc servo loop and squeezed-inverter stage for power-
efficient biopotential sensing,” Sensors, vol. 20, no. 7, p. 2059, 2020.
[71] R. R. Harrison and C. Charles, “A low-power low-noise cmos amplifier for neural
recording applications,” IEEE Journal of solid-state circuits, vol. 38, no. 6, pp. 958–
965, 2003.
[72] N. Laskar, K. Guha, S. Nath, S. Chanda, K. Baishnab, P. Paul, and K. Rao, “Design
of high gain, high bandwidth neural amplifier ic considering noise-power trade-off,”
Microsystem technologies, vol. 27, no. 2, pp. 585–599, 2021.
[73] H. S. Bindra, C. E. Lokin, D. Schinkel, A.-J. Annema, and B. Nauta, “A 1.2-v
dynamic bias latch-type comparator in 65-nm cmos with 0.4-mv input noise,” IEEE
journal of solid-state circuits, vol. 53, no. 7, pp. 1902–1912, 2018.
[74] R. Wu, J. H. Huijsing, and K. A. Makinwa, Precision instrumentation amplifiers and
read-out integrated circuits. Springer Science & Business Media, 2012.
[75] K. Abdelhalim and R. Genov, “Compact chopper-stabilized neural amplifier with low-
distortion high-pass filter in 0.13 µm cmos,” in 2012 IEEE International Symposium
on Circuits and Systems (ISCAS). IEEE, 2012, pp. 1075–1078.
[76] A. Samiei and H. Hashemi, “A chopper stabilized, current feedback, neural recording
amplifier,” IEEE solid-state circuits letters, vol. 2, no. 3, pp. 17–20, 2019.
[77] J. Wu, M.-K. Law, P.-I. Mak, and R. P. Martins, “A 2µw 45nv/Hz readout front
end with multiple-chopping active-high-pass ripple reduction loop and pseudofeed-
back dc servo loop,” IEEE Transactions on Circuits and Systems II: Express Briefs,
vol. 63, no. 4, pp. 351–355, 2015.
[78] A. Nikas, S. Jambunathan, L. Klein, M. Voelker, and M. Ortmanns, “A continuous-
time delta-sigma modulator using a modified instrumentation amplifier and current
reuse dac for neural recording,” IEEE Journal of Solid-State Circuits, vol. 54, no. 10,
pp. 2879–2891, 2019.
108 BIBLIOGRAPHY
[79] S. Wang, M. Ballini, X. Yang, C. Sawigun, J.-W. Weijers, D. Biswas,
N. Van Helleputte, and C. M. Lopez, “A compact chopper stabilized δ-δ σ neural
readout ic with input impedance boosting,” IEEE Open Journal of the Solid-State
Circuits Society, vol. 1, pp. 67–78, 2021.
[80] R. C. Nongpiur, D. J. Shpak, and A. Antoniou, “Improved design method for nearly
linear-phase iir filters using constrained optimization,” IEEE Transactions on Signal
Processing, vol. 61, no. 4, pp. 895–906, 2012.
[81] C. Gold, D. A. Henze, C. Koch, and G. Buzsaki, “On the origin of the extracellular
action potential waveform: a modeling study,” Journal of neurophysiology, vol. 95,
no. 5, pp. 3113–3128, 2006.
[82] R. Quian Quiroga and S. Panzeri, “Extracting information from neuronal popula-
tions: information theory and decoding approaches,” Nature Reviews Neuroscience,
vol. 10, no. 3, pp. 173–185, 2009.
[83] M. S. Lewicki, “A review of methods for spike sorting: the detection and classification
of neural action potentials,” Network: Computation in Neural Systems, vol. 9, no. 4,
p. R53, 1998.
[84] E. M. Schmidt, “Computer separation of multi-unit neuroelectric data: a review,”
Journal of neuroscience methods, vol. 12, no. 2, pp. 95–111, 1984.
[85] F. Kloosterman, S. P. Layton, Z. Chen, and M. A. Wilson, “Bayesian decoding using
unsorted spikes in the rat hippocampus,” Journal of neurophysiology, 2014.
[86] A. Eftekhar, E. P. Sivylla, and G. C. Timothy, “Towards a next generation neu-
ral interface: Optimizing power, bandwidth and data quality,” in 2010 Biomedical
Circuits and Systems Conference (BioCAS). IEEE, 2010, pp. 122–125.
[87] D. Wendler, D. De Dorigo, M. Amayreh, A. Bleitner, M. Marx, R. Willaredt, and
Y. Manoli, “A 0.0046-mm2two-step incremental delta–sigma analog-to-digital con-
verter neuronal recording front end with 120-mvpp offset compensation,” IEEE Jour-
nal of Solid-State Circuits, 2022.
[88] C. Lee, T. Jeon, M. Jang, S. Park, J. Kim, J. Lim, J.-H. Ahn, Y. Huh, and Y. Chae,
“A 6.5-µw 10-khz bw 80.4-db sndr g m-c-based ct modulator with a feedback-assisted
g m linearization for artifact-tolerant neural recording,” IEEE Journal of Solid-State
Circuits, vol. 55, no. 11, pp. 2889–2901, 2020.
[89] B. Gosselin, A. E. Ayoub, J.-F. Roy, M. Sawan, F. Lepore, A. Chaudhuri, and
D. Guitton, “A mixed-signal multichip neural recording interface with bandwidth
reduction,” IEEE Transactions on Biomedical Circuits and Systems, vol. 3, no. 3,
pp. 129–141, 2009.
BIBLIOGRAPHY 109
[90] G. Foffani, M. Morales-Botello, and J. Aguilar, “Spike timing, spike count, and tem-
poral information for the discrimination of tactile stimuli in the rat ventrobasal com-
plex,” Journal of Neuroscience, vol. 29, no. 18, pp. 5964–5973, 2009.
[91] R. Rieger, J. Taylor, A. Demosthenous, N. Donaldson, and P. J. Langlois, “Design of
a low-noise preamplifier for nerve cuff electrode recording,” IEEE Journal of Solid-
State Circuits, vol. 38, no. 8, pp. 1373–1379, 2003.
[92] S. M. Sharroush, Y. S. Abdalla, A. A. Dessouki, and E. A. El-Badawy, “Subthreshold
mosfet transistor amplifier operation,” in 2009 4th International Design and Test
Workshop (IDT), 2009, pp. 1–6.
[93] K. G. Lamb, S. J. Sanchez, and W. T. Holman, “A low noise operational amplifier
design using subthreshold operation,” in Proceedings of 40th Midwest Symposium on
Circuits and Systems. Dedicated to the Memory of Professor Mac Van Valkenburg,
vol. 1. IEEE, 1997, pp. 35–38.
[94] R. Harrison, “A low-power integrated circuit for adaptive detection of action poten-
tials in noisy signals,” Proceedings of the 25th Annual International Conference of
the IEEE Engineering in Medicine and Biology Society (IEEE Cat. No.03CH37439),
vol. 4, p. 3325–3328, 2003.
[95] K. Kundert, “Simulating switched-capacitor filters with spectrerf,” The Designer’s
Guide Community, pp. 1–25, 2006.
[96] T. Sepke, P. Holloway, C. G. Sodini, and H.-S. Lee, “Noise analysis for comparator-
based circuits,” IEEE Transactions on Circuits and Systems I: Regular Papers,
vol. 56, no. 3, pp. 541–553, 2008.
[97] T. Jochum, T. Denison, and P. Wolf, “Integrated circuit amplifiers for multi-electrode
intracortical recording,” Journal of neural engineering, vol. 6, no. 1, p. 012001, 2009.
[98] A. Serb, J. Bill, A. Khiat, R. Berdan, R. Legenstein, and T. Prodromakis, “Unsu-
pervised learning in probabilistic neural networks with multi-state metal-oxide mem-
ristive synapses,” Nature communications, vol. 7, no. 1, p. 12611, 2016.
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Recent advances in multi-electrode array technology have made it possible to monitor large neuronal ensembles at cellular resolution in animal models. In humans, however, current approaches restrict recordings to a few neurons per penetrating electrode or combine the signals of thousands of neurons in local field potential (LFP) recordings. Here we describe a new probe variant and set of techniques that enable simultaneous recording from over 200 well-isolated cortical single units in human participants during intraoperative neurosurgical procedures using silicon Neuropixels probes. We characterized a diversity of extracellular waveforms with eight separable single-unit classes, with differing firing rates, locations along the length of the electrode array, waveform spatial spread and modulation by LFP events such as inter-ictal discharges and burst suppression. Although some challenges remain in creating a turnkey recording system, high-density silicon arrays provide a path for studying human-specific cognitive processes and their dysfunction at unprecedented spatiotemporal resolution. Neuropixels probes were used to simultaneously record from more than 200 cortical neurons in human participants during neurosurgical procedures. The approach could reveal insights underlying human cognition and pathology.
Article
Full-text available
This paper presents a scalable neural recording analog front-end architecture enabling simultaneous acquisition of action potentials, local field potentials, electrode DC offsets and stimulation artifacts without saturation. By combining a DC-coupled Δ\Delta - ΔΣ\Delta \Sigma architecture with new bootstrapping and chopping schemes, the proposed readout IC achieves an area of 0.0077 mm 2 per channel, an input-referred noise of 5.53 ± 0.36 μVrms\mu \text{V}_{\mathrm{ rms}} in the action potential band and 2.88 ± 0.18 μVrms\mu \text{V}_{\mathrm{ rms}} in the local field potential band, a dynamic range of 77 dB, an electrode-DC-offset tolerance of ±70 mV and an input impedance of 663 MΩ\text{M}\Omega . To validate this neural readout architecture, we fabricated a 16-channel proof of-concept IC and validated it in an in vitro setting, demonstrating the capability to record extracellular signals even when using small, high-impedance electrodes. Because of the small area achieved, this architecture can be used to implement ultra-high-density neural probes for large-scale electrophysiology.
Article
Full-text available
Recording reliably extracellular neural activities is an essential prerequisite for the development of bioelectronics and neuroprosthetic applications. Recently, a fully differential, two-stage, integrating pre-amplifier was proposed for amplifying and then digitising neural signals. The amplifier featured a finely tuneable offset that was used as a variable threshold detector. Given that the amplifier is integrating, the DC operating point keeps changing during integration, rendering traditional analysis (AC/DC) unsuitable. In this work, we analyse the operation of this circuit and propose alternative definitions for validating the necessary key performance metrics, including gain, bandwidth, offset tuning range and offset sensitivity with respect to the memory states of the employed memristors. The amplification process is analysed largely through investigating the transient behaviour during the integration phase. This benchmarking approach is finally leveraged for providing useful insights and design trade-offs of the memristor-based integrating amplifier.
Article
This article presents a recording front end for high-density CMOS neuronal probes with in situ digitization and electrode offset voltage compensation. The analog front end (AFE) is based on a continuous-time (CT) two-step (TS) incremental delta–sigma ( IΔΣ\text{I}\Delta \Sigma ) analog-to-digital converter (ADC) with an extended counting technique and features an input offset voltage compensation of 120 mVpp. Hardware sharing in the TS quantization process allows the integration of the front end in an area of only 0.0046 mm 2 and, thus, directly under the recording electrodes on the shank of a probe in 180-nm CMOS. The average integrated noise is as low as 4.88 μVrms\mu \text{V}_{\text {rms}} , 4.46 μVrms\mu \text{V}_{\text {rms}} , and 2.51 μVrms\mu \text{V}_{\text {rms}} in the full bandwidth of 0 Hz–10 kHz, in the frequency band of action potentials (AP, 0.3–10 kHz) and local field potentials (LFP, 0.5 Hz–1 kHz), respectively. Each recording front end consumes 8.57 μW\mu \text{W} , and transmitting the digitized data to an external host needs additionally 6.05 μW\mu \text{W} per channel.
Article
Memristive technology has experienced explosive growth in the last decade, with multiple device structures being developed for a wide range of applications. However, transitioning the technology from the lab into the marketplace requires the development of an accessible and user-friendly design flow, supported by an industry-grade toolchain. In this work, we demonstrate the behaviour of our in-house fabricated custom memristor model and its integration into the Cadence Electronic Design Automation (EDA) tools for verification. Various input stimuli were given to record the memristive device characteristics both at the device level as well as the schematic level for verification of the memristor model. This design flow from device to industrial level EDA tools is the first step before the model can be used and integrated with Complementary Metal-Oxide Semiconductor (CMOS) in applications for hybrid memristor/CMOS system design.
Article
\normalsize The capability of in-memory computation, reconfigurability, low power operation as well as multistate operation of the memristive device deems them a suitable candidate for designing electronic circuits with a broad range of applications. Besides, the integrability of memristor with CMOS enables it to use in logic circuits too. In this work, we demonstrate with examples the design flow for memristor-based electronics, after the custom memristor model already being integrated and validated into our chosen Computer-Aided Design (CAD) tool to performing layout-versus-schematic and post-layout checks including the memristive device. We envisage that this step-by-step guide to introducing memristor into the standard integrated circuit design flow will be a useful reference document for both device developers who wish to benchmark their technologies and circuit designers who wish to experiment with memristive-enhanced systems.
Article
This article introduces a compact NMOS-only voltage reference that is able to operate down to a 0.25-V supply voltage and 5.4-pW power consumption. This allows reliable generation of a stable output voltage even in harvested systems under highly uncertain environmental conditions. At the system level, pW-power, 0.25-1.8-V operation, and a competitive power supply rejection ratio (PSRR) relax or eliminate the need for intermediate power conversion and supply regulation for additional power and cost reductions. The proposed voltage reference is based on a body biasing scheme assisted by replica well biasing to compensate voltage and temperature fluctuations. A 180-nm test chip shows that the proposed reference occupies an area of 2200 μm², while providing 91.4-mV output voltage with 0.51-mV (0.56%) standard deviation, 24.2 μV/°C (265 ppm/°C) temperature coefficient, and 144.5 μV/V (0.16%/V) line sensitivity across 30 dice from the same manufacturing lot. The resulting absolute output voltage accuracy without any trimming is 2.8 mV at 3-σ variations, 0.3-V fluctuation, and temperature change by 50 °C, improving on prior art by 1.4-19.7x. Overall, the capability of reliable operating down to ultralow voltages, pW-power, and the inherently small footprint make the proposed reference well suited for low-cost tightly constrained systems.