Next Article in Journal
The Conceptual Design of a Mechatronic System to Handle Bedridden Elderly Individuals
Previous Article in Journal
Color Restoration of RGBN Multispectral Filter Array Sensor Images Based on Spectral Decomposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Low-Noise Transimpedance Amplifier for BLM-Based Ion Channel Recording

by
Marco Crescentini
1,2,*,†,
Marco Bennati
3,†,
Shimul Chandra Saha
4,†,
Josip Ivica
4,†,
Maurits De Planque
4,†,
Hywel Morgan
4,† and
Marco Tartagni
1,2,†
1
Department of Electrical Electronic and Information Engineering G. Marconi, University of Bologna, Cesena Campus, Via Venezia 52, IT-47521 Cesena, Italy
2
Advanced Research Center on Electronic Systems (ARCES), University of Bologna, Cesena Campus, Via Venezia 52, IT-47521 Cesena, Italy
3
Center for Industrial Research (CIRI-ICT), University of Bologna, Via Venezia 52, IT-47521 Cesena, Italy
4
Department of Electronic and Computer Science & Institute for Life Science, University of Southampton, University Road, SO17 1BJ Southampton, UK
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Sensors 2016, 16(5), 709; https://doi.org/10.3390/s16050709
Submission received: 11 December 2015 / Revised: 13 April 2016 / Accepted: 5 May 2016 / Published: 19 May 2016
(This article belongs to the Section Biosensors)

Abstract

:
High-throughput screening (HTS) using ion channel recording is a powerful drug discovery technique in pharmacology. Ion channel recording with planar bilayer lipid membranes (BLM) is scalable and has very high sensitivity. A HTS system based on BLM ion channel recording faces three main challenges: (i) design of scalable microfluidic devices; (ii) design of compact ultra-low-noise transimpedance amplifiers able to detect currents in the pA range with bandwidth >10 kHz; (iii) design of compact, robust and scalable systems that integrate these two elements. This paper presents a low-noise transimpedance amplifier with integrated A/D conversion realized in CMOS 0.35 μm technology. The CMOS amplifier acquires currents in the range ±200 pA and ±20 nA, with 100 kHz bandwidth while dissipating 41 mW. An integrated digital offset compensation loop balances any voltage offsets from Ag/AgCl electrodes. The measured open-input input-referred noise current is as low as 4 fA/√Hz at ±200 pA range. The current amplifier is embedded in an integrated platform, together with a microfluidic device, for current recording from ion channels. Gramicidin-A, α-haemolysin and KcsA potassium channels have been used to prove both the platform and the current-to-digital converter.

1. Introduction

Ion channels are nanoscale pores that sit in the cell membrane, allowing communication of the cell with the external environment through ionic currents. The open/close behavior of ion channel is modulated through different mechanisms, e.g., voltage, ligand binding, pH change, or mechanical strain. Channels are crucial for the control of physiology and any malfunction is at the root of a variety of pathologies and diseases [1]. Ion channel recording is an important component of the next generation HTS diagnostic tools used for drug discovery, DNA sequencing and single molecule detection [2]. There are two main techniques for ion channel screening:
  • Patch clamp, where a glass pipette, or a micro-aperture in a solid-state device is used to pull a patch of cell membrane [1,3,4] (Figure 1).
  • Planar bilayer lipid membranes (BLM), where a single ion channel is inserted into a lipid bilayer suspended over a micro-aperture [5] (Figure 1).
The patch-clamp technique is widely used in modern HTS instruments. The advantages of patch-clamp are high fidelity, since the ion channels exist in their native physiological environment, together with a high level of automation and parallelization [4,6]. This technique suffers from low specificity and high noise since a number of different ion channels are measured together, and also the membrane provides a large capacitance. On the contrary, BLM technique provide excellent electrical sealing and high sensitivity detection, down to single molecule, with minimum noise and capacitance [7,8,9]. The design of HTS system based on BLM ion channel recording faces three main challenges [10]:
  • A microfluidic device allowing stable, reliable and automatic BLM formation.
  • A fast low-noise electronic interface ables to acquire pA currents.
  • A compact, robust and scalable system containing an array of microfluidic devices and electronic interface.
This paper focuses on the second challenge above; the other two challenges have been discussed previously [11,12,13]. The electronic readout is a key element in the design of a BLM-based HTS system. The main requirements for the electronic interface are low-noise (noise floor <10 fA/√Hz), high-sensitivity (transresistance >1 GΩ) and wide-bandwidth (>10 kHz) [14,15]. Specific requirements are mainly related to the kind of ion channel under investigation. For instance, potassium ion channels, such as KcsA, have fast responses (100 μs) and zero-voltage conductivity lower than 100 pS, resulting in currents of the order of a few pA with applied voltages lower than 100 mV [16]. In general an ion channel has (i) very high output impedance (from 1 to 100 GΩ); (ii) noise level smaller than 1 pArms at 1 kHz; (iii) open/close events ranging from few milli-seconds to hundreds of micro-seconds and (iv) capacitance of the order of tens of pF [10,15,17].
The benchmark for low-noise low-current recording is the Axon Axopatch 200B, which has 100 kHz bandwidth and input-referred noise of 6 fA/√Hz in resistive mode and of 0.7 fA/√Hz in capacitive mode, but it is a bulky instrument not suitable for parallel recording [18]. A great number of low-noise low-current readout circuits have been presented in the literature in the last few years, but none of them completely fits the requirements. Hsu et al. [19] presented two different designs both achieving 5 fA/√Hz (160 fArms at 1 kHz) but with different weaknesses: one has 560 kHz bandwidth but an insufficient gain of 100 MΩ; while the other has enough gain (4.7 GΩ) over a narrow bandwidth of 1 kHz. Moreover, both the circuits are realized using discrete components, so they are not the best solutions when highly-parallel (>1024 channels) HTS systems have to be designed. Jafari et al. [20], as well as Crescentini et al. [13], presented very low-noise CMOS frontends with high gain (>1 GΩ) and noise floor as low as 2 fA/√Hz (63 fArms at 1 kHz) and 3 fA/√Hz (95 fArms at 1 kHz) respectively, but they are limited in acquisition bandwidth, which was lower than 10 kHz. Rosenstein et al. [21] described a fast current readout IC for high-throughput DNA sequencing; the circuit has more than 1 MHz bandwidth but the noise floor is limited to 12 fA/√Hz (380 fArms at 1 kHz).
This paper presents a low-noise transimpedance amplifier realized in CMOS 0.35 μm technology with a measured input-referred noise as low as 4 fA/√Hz (133 fArms at 1 kHz), a gain of 2.25 GΩ and 100 kHz bandwidth. The transimpedance amplifier is based on integrator-differentiator scheme [14]. The CMOS implementation is scalable in terms of the number of concurrently acquired channels while minimizing the stray input capacitance and interference, with benefits in the noise performance since the noise is linked to the input capacitance [13,14]. An integrator-differentiator scheme provides a current sensing interface with the lowest noise floor, but suffers from saturation of the integrator stage [14]. To avoid saturation while maintaining a wide acquisition bandwidth and limiting the noise sources, we propose a periodic reset of the readout circuits at frequency fR with A/D sampling at frequency fS >> fR, disregarding the reset behavior. In this way the folding noise due to sampling is reduced and the bandwidth is not limited by the reset. A second-order delta-sigma (ΔΣ) analog-to-digital converter (ADC) oversamples the signal at 10 MHz and generates a 1-bit 10 MS/s digital stream that is decimated by digital FIR filter implemented on a FPGA. This solution simplifies the signal routing when concurrently acquiring a great number of channels, and gives a flexible bandwidth-noise trade-off to the user by acting on the oversampling ratio (OSR) parameter in the decimator filter [22]. The system also integrates a digital offset cancellation loop (OCL) balancing any voltage offset from Ag/AgCl electrodes. The amplifier has been validated, together with microfluidic devices by measuring the activity of three different ion channels: gramicidin-A, α-haemolysin and KcsA potassium channels.
Section 2 briefly presents the overall platform and the microfluidic devices then describes the implementation of the CMOS transimpedance amplifier circuit with detailed noise analysis. Finally, Section 3 reports experimental measurements and validation of the proposed readout circuit.

2. Proposed Transimpedance Amplifier

2.1. Ion Channel Recording Platform

The complete ion channel recording platform is able to concurrently acquire 12-channels, and is composed of (Figure 2):
  • Three disposable microfluidic devices manufactured on a glass substrate holding 4 BLMs each [12].
  • A small PCB hosting two CMOS 2-channel low-noise current-to-digital amplifiers that can measure pA currents.
  • A motherboard with a digital control unit implemented in a Field Programmable Gate Array (FPGA) [11].
All the components are integrated onto a single platform, offering a fully scalable acquisition system. The system architecture was presented for the first time in [23], while the ability to concurrently acquire multiple channels was previously published in [11] and [13]. This paper focuses on the design rationale of the analog frontend of the CMOS current-to-digital amplifier. For a description of the parallel microfluidic platform refer to [11,12,13].
A block diagram of the platform is shown in Figure 3. The functionality of the system is as follows. The CMOS IC applies a voltage stimulus VSTIM to the ion channel through the low-noise amplifier (LNA) virtual short circuit; this stimulus could be either a constant voltage or a time-varying voltage. The ionic current flowing through the ion channel is translated into an electronic current by Ag/AgCl electrodes in the microfluidic device. The CMOS IC acquires the input current IIN and digitizes it into a 1-bit data stream. It uses a novel scheme for the transimpedance amplifier and a 2nd order ΔΣ modulator targeting 16-bit resolution for the A/D conversion. The analog-to-digital converter (ADC) output is filtered and decimated by a FIR filter implemented on a FPGA. Finally, PC communication is via a USB link.
An internal digital loop compensates for any input voltage offset from the Ag/AgCl electrodes (Voff,ele in Figure 3) [13]. This offset cancellation is done at the beginning of each experiment as follow:
  • Read the front-end output voltage VOUT;
  • Compare VOUT with the reference voltage VCM;
  • Change DC voltage VOFF so that it becomes equal to VCM + Voff,ele. (Note reference electrode is tight to VCM).
Final voltage VC applied to the LNA positive input is given by VC = VSTIM + VOFF so that VOFF counteracts the electrode offset, while VSTIM appears as the voltage drop across bilayer membrane. The system is fully programmable via SPI. Two input ranges are implemented (±20 nA and ±200 pA) with maximum acquisition bandwidths of 100 kHz.

2.2. Microfluidic Device

The microfluidic device holds up to four separate BLMs and is manufactured on a glass substrate [12]. It has dimensions of 15 × 15 mm with integrated Ag/AgCl electrodes. Bilayers were formed over apertures of approximately 100 μm diameter (Figure 4) [12,13]. The measured capacitance of a bilayer suspended across a 75 μm diameter aperture is typically 15–30 pF. Complete description of the microfluidic device can be found in [12].

2.3. Sensing Frontend Rationale

The front-end is based on integrator-differentiator scheme offering maximum noise performance due to the input integrator stage [14]. Figure 5 shows a schematic diagram of the complete front end. The direct signal path is composed of a current integrator, a capacitive voltage amplifier, a continuous-time (CT) differentiator outputting a voltage directly proportional to the input current IIN, and a Sallen-Key low-pass filter. Integrator, voltage amplifier and differentiator are periodically reset to avoid saturation, while the Sallen-Key filter holds the output voltage vOUT during reset.
It is possible to discriminate two different phases as shown in Figure 6:
  • Active phase. During this phase v O U T ( t ) = R e q i I N ( t ) where Req is the equivalent trans-resistance of the amplifier, which is given by:
    R e q = C 2 C 4 C 1 C 3 R 4
  • Reset phase. During this phase the output voltage is kept constant while the rest of the circuit reset.
This 2-phase behavior is controlled by signal F3 internally generated from an external 80 MHz clock. To minimize the effect of charge injection, the control signals F1, F2 and F3 are designed to start at the same time but stopping one after the other, and the switches are realized by transmission-gates with dummy elements.
The proposed architecture differs from a standard discrete-time transimpedance amplifier, as defined in [14], since the sampling frequency fS is unrelated to the reset frequency fR = 1/TR; specifically fS is greater than fR. In this way the acquisition bandwidth is not limited by the periodic reset but the noise becomes cyclostationary. Detailed analysis of the effects of cyclostationary properties of noise is discussed in Section 2.7.
Another important limitation on the maximum frequency is given by the bandwidth of the integrator that is almost equal to:
B W I n t e g r a t o r ~ G B W C 1 C S + C P
where GBW is the unity gain bandwidth of the OTA, CS is the capacitance of the microfluidic setup with the BLM, and CP is the parasitic capacitances due to interconnects and input stage of the transimpedance amplifier. The combination of Equations (1) and (2) sets a trade-off on the value of the feedback capacitance C1 that should be small enough to maximize Req and minimize input noise (see Section 2.7), but large enough to speed up the integrator. C1 was set to 1 pF, where GBW = 92 MHz, CP is of the order of a few pF, and CS is expected to be in the range 40–80 pF [12]. With these parameters, a 1 MHz bandwidth of the integrator is obtained. This value is ten times higher than acquisition bandwidth of the entire system and ensures a fast settling of the integrator after reset. Note that now the acquisition bandwidth is not limited by periodical reset but only by bandwidth of the OTA and parasitic capacitances as reported in Equation (2). The reduction of trans-resistance Req due to the chosen value for C1 is compensated by the gain of the voltage amplifier stage placed between integrator and differentiator.
Timing characteristics of the reset phase, which are duration τR and period TR, affect both the signal and noise. During the reset phase, the output voltage is disconnected from the input, hence the system does not see what the input current actually is, which leads to loss of information. As a result, τR must be minimized while TR must be maximized. The same conclusion comes from noise analysis (see Section 2.7). Note that the reset period TR has an upper limit given by saturation of the first two OTAs. Assuming a maximum 200 pA DC input current (∆IIN) flowing through the input, then the system saturates after a time TSAT given by:
T S A T = C 1 Δ V O 1 Δ I I N
where ∆VO1 is the maximum allowed voltage swing at the integrator output, which is equal to:
Δ V O 1 = Δ V O 2 C 3 C 2 = Δ V O U T C 3 C 2
where we assumed ∆VO2 = ∆VOUT since at low frequency the gain is set by the first two stages. Therefore the reset period T should be:
T T S A T = C 1 C 3 C 2 Δ V O U T Δ I I N = C 1 C 3 C 2 C 2 C 4 C 1 C 3 R 4 = C 4 R 4
where we assumed a maximized full-scale, that is:
Δ V O U T = R e q Δ I I N
Equation (5) shows how the reset period TR is linked to the time constant C4R4; thus C4 and R4 must be maximized. Once C1, C4 and R4 are chosen, the ratio C2/C3 is given by the combination of Equations (5) and (6). The resistor R3, and capacitor C4, creates a first-order low-pass filter, reducing the noise before differentiation.
A list of circuit parameters is reported in Table 1. The output voltage full-scale ∆VOUT is fixed at ±450 mV by the OTA, thus Req = 2.25 GΩ for ∆IIN = ±200 pA and Req = 22.5 MΩ for ∆IIN = ±20 nA.

2.4. ADC

The ADC is integrated in the CMOS current-to-digital converter as shown in Figure 3. The scheme of the ADC is standard and it is reported in Figure 7; it is a switched-capacitors second-order ΔΣ converter operating between voltages VREF+ = 2.1 V and VREF−= 1.2 V. Timing signals, ph1 and ph2, are two non-overlapping 10 MHz signals generated from an external 80 MHz clock. The use of a second-order ΔΣ converter allows keeping the quantization noise below the thermal noise of the input front-end thanks to ΔΣ noise shaping. The output of the ADC is a 10 MHz 1-bit data stream that is filtered and downsampled by a FIR digital filter implemented in the FPGA.

2.5. Stimulus Generation and Offset Compensation

The offset potential arising from the electrode-electrolyte interface can be of the order of tens to hundreds of millivolts, following the Nernst equation [24]. The dispersion of actual value of the offset potential around the nominal value strongly depends on type of the electrode and fabrication technique. Moreover, microfabricated electrodes suffer from higher instability and dispersion due to small sizes of the electrode and low control of process parameters [24,25,26].
This offset potential generates offset current that limits the acquisition range or even causes saturation of the amplifier, since the equivalent transresistance is very high. For instance; 200 mV offset over 1 GΩ channel resistance leads to 200 pA current that saturates the transimpedance amplifier when working in the 200 pA range. To cope with this, the OCL is activated at the beginning of each experiment [27]. It compares VOUT with the reference voltage VCM, which is the bias voltage of the reference electrode, and generates a DC voltage VOFF that is applied to the positive input of the integrator:
V O F F = V C M ( 1 + R e q R S ) + R e q R S V o f f , e l e ( 1 + R e q R S ) V C M + V o f f , e l e
where RS is the equivalent resistance of the ion channel. Under the assumption Req >> RS then the offset current is almost nulled, while a small offset current still remains in case of higher value of RS. Note that offset of the LNA is compensated along with offset of the electrodes. The compensation loop is realized by means of a comparator, a latch and an 8-bit up/down counter working at 150 Hz (Figure 8).
A time-varying stimulus signal vSTIM is needed when working with voltage gated ion channels or for full characterization of ion channels. The signal vSTIM is digitally generated by the FPGA and then sent to the CMOS amplifier through SPI interface, as shown in Figure 8. Both offset compensation and stimulus generation are addressed in the digital domain. Voltages vSTIM and VOFF are added together and converted in analog domain by a 10-bit DAC to create the voltage vC that is applied to the positive input of the integrator (Figure 3):
v C = v S T I M + V O F F
The 10 bits of the DAC must accommodate the swing for both vSTIM and VOFF, limiting them to ±384 mV and ±128 mV, respectively. At the DAC output, a passive LPF filters out the high frequency noise. Extremely low noise acquisitions require an external capacitor of at least 1 nF. Note that a larger external capacitance reduces the noise as well as the bandwidth of the stimulus voltage. For instance, setting CEXT = 1 nF limits the bandwidth of vSTIM to a few kHz.

2.6. Subtractor

The voltage vC = vSTIM + VOFF is applied to the DUT by means of the virtual short circuit imposed by the negative feedback of the integrator. Hence the integrator behaves like a non-inverting amplifier from the vC standpoint, and the voltage v01 can be written as:
v 01 = ( C S + C P C 1 + 1 ) d v C d t d t = i I N C 1 d t + v C
Equation (9) states that vC signal directly propagates through the first stage as an unwanted additive component to the measured input signal. To avoid this effect, the subtractor stage multiplies vC by −1 and adds its output to integrator output (Figure 5). Obviously this stage adds noise but it is not needed for electrophysiology experiments requiring constant vC; hence, it is possible to activate or deactivate it using control signal Sub.

2.7. Noise Analysis

Noise models presented in [14] are not directly applicable to the proposed architecture since it is based on a combination of both continuous time (CT) and discrete time (DT) approaches. Hence a new model is derived, based on theory and methods described in [28,29]. The analog frontend can be simplified as shown in Figure 5a, where the following assumptions were made:
  • all the stages prior to the sampling are treated as linear time-invariant systems;
  • node x takes into account low-pass filtering done by the Sallen-Key but not the sampling; there is not a direct correspondence of node x in the schematic diagram (Figure 5b).
  • node OUT is renamed into y to get more compact equations.
The noise power spectrum density (PSD) at node x can be written as:
G x ( f ) ~ ( 2 π f ) 2 ( C 1 + C I N ) 2 R e q 2 e n 2 ¯ 1 ( 1 + f f p ) 2
where CIN = CS + CP, en is the input-referred noise source of the OTA, fp is the dominant pole of the system that is set by the LPF, and noises generated by voltage amplifier and differentiator have been neglected. Note that Gx(f) has not a white shape but it rises with f where Flicker dominates, and with f2 where thermal noise dominates (Figure 9a). Simplifying Gx(f) to a triangular shape the autocorrelation function of x(t) becomes:
R x x ( τ ) ~ A f p sinc 2 ( f p τ )
where A is the peak value of Gx(f). The voltage at node y can be seen as v y ( t ) = y ( t ) + y ( t ) :
y ( t ) = { v x ( t )   in active phase 0   in reset phase  y = v x ( t ) n p ( t n T )   y ( t ) = { 0 in active phase v x ( n T ) in reset phase    y ( t ) = n v x ( n T T R ) [ 1 p ( t n T ) ]
where p(t) is a step function equal to zero in the reset phase and equal to 1 for every other time. Signal y′(t) takes into account the linear CT behavior of the system, although periodicity creates cyclostationary properties [30], while y″(t) takes into care the DT sampling behavior. The noise PSD at node y can be written as [31]:
G y ( f ) = G y ( f ) + G y ( f ) + 2 Re { F [ R y y ( t , τ ) T R ] }
where F denotes the Fourier transform, Ry′y″ is the cross-correlation function and TR is the reset period.
The first term in Equation (13) can be expanded as:
G y ( f ) = 1 T R 2 n ( T R τ R ) 2 sinc 2 [ n ( T R τ R ) T R ] G x ( f n T R )
where τR is the pulse duration of F3. Note that periodicity of y′ leads to folding of the noise PSD. However the sinc function is around zero for all n0 since τR << TR; hence Equation (14) can be simplified to:
G y ( f ) = ( T R τ R ) 2 T R 2 G x ( f ) G x ( f )
This simplification is even more valid when τR tends to zero (i.e., pure CT behavior) while the whole term in Equation (14) goes to zero when τR tends to TR (i.e., pure DT behavior).
The second term in Equation (13) is the noise PSD of a standard sampling process, hence it can be written as [14,31,32]:
G y ( f ) = τ R 2 T R 2 sinc 2 ( f τ R ) k G x ( f k T R )
The third term in Equation (13) needs a little more derivation. Denoting random variables with upper case letters, then the cross-correlation function can be written as:
R y y ( t , τ ) = E [ Y ( t ) Y ( t ) ] = E [ X ( t ) n p ( t n T R ) m X ( m T R τ R ) [ 1 p ( t τ m T R ) ] ] = R x x ( τ R ) n m δ ( t m T R ) p ( t n T R ) [ 1 p ( t τ m T R ) ]
Since Rxx(τ) has the form expressed in Equation (11), then RxxR) is around zero for τR > 1/fp. In our case τR = 4.8 μs and fp is around 710 kHz, hence we neglect the cross-correlation term in Equation (13). If 1/fp > τR this simplification is not valid any longer and Equation (17) should be considered.
Using all the above, we simplify Equation (13) to:
G y ( f ) = G x ( f ) + ( τ R T R ) 2 sinc 2 ( f τ R ) k G x ( f k T R )
Note that high frequency thermal noise gives the main contribution to folding noise while Flicker noise can be easily ignored in both the terms because of its frequency shape (see Figure 9b). The first term in Equation (18) is the noise PSD before sampling, while the second term in Equation (18) is the folding of high frequency noise due to the reset process. Unfortunately the folding term cannot be easily simplified using the undersampling ratio (USR) as done in [14,28], because Gx(f) has not a standard white shape. Finally the input-referred noise is given by:
i i n 2 ¯ ( f ) = G y ( f ) R e q 2 G x ( f ) + ( τ R T R ) 2 sinc 2 ( f τ R ) k = 0 U S R G x ( f k T R ) R e q 2
Equation (19) was implemented in Matlab by solving the summation for k up to USR = πfPTR. Matlab analysis reveals that folding noise dominates the first term in Equation (19), although it is multiplied by a pre-factor (τR/TR)2 that is much less than 1. Note that this pre-factor is near 1 in pure-DT approaches.
Equation (19) indicates some design considerations:
-
The most direct method of reducing folding noise is lowering the USR, which defines how many times the noise folds back into the baseband. This can be easily done by lowering fp, but this directly affects the bandwidth of the system and the sampling error [14]. Moreover, if 1/fp becomes greater than the reset pulse duration τR then Equations (18) and (19) are no longer valid, since cross-correlation power computed from Equation (17) must be taken into consideration.
-
Another important parameter is the period TR, which appears in both USR and pre-factor. It should be small to lower USR while it should be big to lower the pre-factor (τR/TR)2. Since TR is squared in the pre-factor term then it is better to make it as high as possible. This relation between noise and parameter TR is confirmed by periodic-steady-state noise (pss-noise) analysis and it is predicted by our mathematical model, see Figure 10.
-
Another way of reducing the folding noise is to keep Gx(f) as low as possible. This directly translates into using a low-noise OTA and lowering the input capacitance CIN as well as the feedback capacitance C1, creating a noise-bandwidth trade-off.

3. Experimental Results

3.1. Implementation

The system has been implemented in 0.35 μm CMOS technology. A microphotograph of the ASIC is shown in Figure 11. It has two separate current frontends occupying a total area of 9.2 mm2. The two-core design was dictated by the need to implement a prototype compact parallel recording platform, with 12 channels acquired concurrently [13]. A single core consumes a total of 41 mW. This power consumption includes all the circuits implemented in the CMOS chip, i.e., analog frontend, ADC, DAC, voltage references, clock buffering and digital circuits. The system can acquire signals up to 100 kHz at the highest resolution.

3.2. Noise Measurements

The −3 dB bandwidth reported in following figures is set by the digital FIR filters at the output of the ΔΣ ADC. We implemented 150 taps FIR filters with triangular windowing at different cut-off (−6 dB) frequencies (10 kHz and 200 kHz). The −3 dB bandwidth of these filters is 7.5 kHz and 175 kHz respectively. Note that the acquisition bandwidth can be theoretically increased up to half the ADC sampling frequency (i.e., 5 MHz), losing in resolution. However, the Sallen-Key LPF in Figure 5 and the quantization noise practically limit the maximum bandwidth to 710 kHz and 200 kHz, respectively. For bandwidths higher than 200 kHz, i.e., OSR lower than 26, the quantization noise dominates over the thermal noise reported hereafter.
The input-referred noise PSD is computed by dividing the output noise PSD by the square of the equivalent transresistance Req. Figure 12a shows the open-input input-referred noise measured at both the 200 pA and 20 nA input ranges with deactivated subtractor stage. The system has an input noise as low as 4 fA/√Hz in the 7.5 kHz bandwidth, and 6 fA/√Hz in the 175 kHz bandwidth, proving the low-noise capability and the wide acquisition bandwidth. Noise PSDs are flat at low frequencies since folding noise dominates, as discussed in Section 2.7, while they rise at high frequencies where the CT term dominates, as usual in transimpedance amplifiers [14].
Table 2 compares the noise model given by Equation (19) with measured and simulated r.m.s. noise over a 10 kHz bandwidth. The theoretical value was obtained by implementing Equation (19) in Matlab and solving the summation for k up to USR = πfPT. The parameters are the followings: fp = 710 kHz, TR = 102.4 μs, τR = 4.8 μs, C1 = 1 pF, Req = 2.25 GΩ, en = 3 nV/√Hz. Noise simulation was done using SpectreRF®, which takes into account noise folding and cyclostationary properties of the system. Note that CIN = CP = 3 pF was used in both mathematical model and simulations to count for stray capacitances facing to the input node, such as capacitive effects due to pad, bonding wires, pin, etc. This value was indirectly estimated from measurements and parasitic extraction.
Figure 12a also shows the input-referred noise current recorded at maximum bandwidth with the 20 nA input range selected. The noise floor is ten times higher than noise measured at the 200 pA range. Spikes around 10 kHz and multiple frequencies results from the periodicity of the system due to the cyclostationary nature of the output noise [30]. This behavior can be seen as periodic spikes in the time-domain (Figure 12b). The noise is closely dependent on CIN as seen in Equation (10) and demonstrated by measurements shown in Figure 12c. This result confirms the need for small microfluidic device and integrated electronic readout placed as close as possible to the microfluidic chip. All the noise PSDs described above refer to measurements done in the optimum condition of deactivated OCL and constant voltage vC.

3.3. Offset Compensation Loop and Subtractor

Demonstration of the effectiveness of the OCL is reported in Figure 13a. The presence of the electrode offset causes a nA current to flow into the transimpedance amplifier. The OCL changes the DC component of the stimulus voltage so as to counteract the electrode offset and apply a zero DC voltage to the BLM.
Subtractor stage has been tested connecting 10 pF to the input node and applying a 100 mVpp triangular wave as vC signal. Under this condition, a 400 pA (800 pApp) square wave current flows through the input node. The amplifier works at ±20 nA range and 7.5 kHz bandwidth. Figure 13b reports the estimated input current, computed referring the output of the FIR filter back to the input by means of the equivalent transresistance, when subtractor stage is either deactivated or activated. Following the analysis described in section II, the estimated input current is given by:
i ^ I N = v o u t R e q = ( 1 + C 1 C S ) i I N
In the former case, the input current is overestimated by a factor of 2, while the latter case reports the correct 400 pA value, proving the functionality of the subtractor stage. The activation of the OCL and the subtractor adds noise sources in the circuit increasing the input-referred noise, as shown in Figure 13c.

3.4. Ion Channel Recording

The system was tested by acquiring data of single ion-channel currents. The measurement setup consists of a single microfluidic device with its own ASIC frontend mounted on the platform shown in Figure 2. Validation tests were done with three different kinds of ion-channel: gramicidin-A, α-haemolysin and KcsA potassium channel.

3.4.1. Gramicidin-A

The microfluidic device was filled with buffer solution (1 M KCl, 10 mM HEPES, pH 7.4) and bilayers formed by painting a 10 mg/mL solution of phospholipid (1,2-diphytanoyl-sn-glycero-3-phosphocholine, DPhPC, Avanti Polar Lipids, Alabaster, AL, USA) in decane over the apertures. Figure 14a shows current traces for gramicidin-A ion-channels with an applied potential of 100 mV. Data was acquired at 625 Hz, and shows the opening of three independent ion-channels (each current step corresponds to a single gramicidin A dimer). The typical ion channel current step is about 2.6 pA, which corresponds to a conductance of ~26 pS and matches literature values for gramicidin-A [33,34]. This simple test proves the ability of the system to acquire low current signals involved in standard ion-channel monitoring.

3.4.2. α-Haemolysin

To demonstrate the capability of the system to record larger current steps, a 2.5 μg/mL solution of the protein α-HL was added to the top aqueous compartment. Figure 14b shows current steps of the order of 50 pA, indicative of the insertion of single α-HL nanopores, at an applied potential of 50 mV. A small number of current steps is to be expected because α-HL does not exhibit a transition between a closed and an open state. Again, the magnitude of the current steps is in reasonable agreement with literature values for α-HL in 1 M KCl solution [35,36].

3.4.3. KcsA Potassium Channel

To demonstrate fast gating characteristic, the wild type KcsA channel, which has rapid closing and opening events was used. The BLM was made from a mix of POPC and POPG (1:1) lipids, due to the influence of PG on channel gating [16]. KcsA activates only in acidic pH, therefore both the top and bottom compartment was filled with 150 mM KCl, 10 mM HEMS, pH 4.0. A 5 μL suspension of proteoliposomes containing KcsA channels (1000:1 lipid to protein ratio) was added to the top compartments and the potential applied to the bottom compartment. Characteristically, KcsA displays long-lived ‘quiet’ or closed intervals and exhibits low open probabilities [16]. Single channel traces in symmetrical 150 mM K+ solutions, at 120 mV are shown in Figure 14c,d, where the current steps are consistent with previous published results [35]. The data shows the platform can record fast gating channels.
It was not possible to perform ion channel recording at higher bandwidths due to limitations in the microfluidic-biological setup; that are mainly capacitive loading and biological noise. In fact, the channel activity is barely visible even in the 10 kHz-filtered (−6 dB) acquisition of Figure 14d. Increasing the acquisition bandwidth to 100 kHz will lead to a three-fold higher noise, at least.

3.5. State-of-the-Art Comparison

The benchmark instrument for low-noise acquisition of small currents (<1 nA) is the Axon Axopatch 200B (Molecular Devices, Sunnyvale, CA, USA), which has an input-referred noise of 6 fA/√Hz in resistive mode, and 0.7 fA/√Hz in capacitive mode [18]. A comparison between Axopatch and the transimpedance ASIC is shown in Figure 15. Our system has comparable performances both in term of bandwidth and noise but integrated in a single silicon chip.
Comparison with state-of-the-art low-noise CMOS current interfaces is reported in Table 3. Our system has a state-of-the-art noise performance (6 fA/√Hz) together with wide acquisition bandwidth (100 kHz) and high gain (2.25 GΩ), meeting all the requirements of BLM ion-channel recording. It also embeds the ADC simplifying the design of parallel acquisition platform. Note that power consumption is not a main concern for this type of application, providing it is lower than a few Watts. Ferrari et al. [37] achieves optimum noise performance and wide bandwidth but the output is still a current, needing a I/V stage that may introduce extra noise. Moreover, the system is shot-noise limited, showing a noise floor that increases with the input current. Reference [38] is the only CMOS transimpedance amplifier presenting low-noise (<10 fA/√Hz) and wide bandwidth (>100 kHz), but does not embed the ADC.

4. Conclusions

This paper presented a CMOS transimpedance amplifier based on the integrator-differentiator scheme for BLM ion-channel recording. The application requires low-noise current acquisition (~fA/√Hz), high transimpedance value (>1 GΩ) and wide acquisition bandwidth (>10 kHz) in a compact design that can be easily parallelized for a future implementation in HTS instruments. CMOS integration of the architecture is a key feature since parasitic capacitances on the input node strongly increase the electronic noise.
The integrator-differentiator scheme is periodically reset to avoid saturation but the output voltage is sampled at frequency fS > fR disregarding the reset behavior. In this way the circuit achieves wide acquisition bandwidth and low noise performance even at low frequencies. The final result is a current-to-digital converter meeting all the applications requirements: (i) noise floor less than 10 fA/√Hz (i.e., 400 fArms at 10 kHz and 1.9 pArms at 100 kHz, both measured at ±200 pA range); (ii) 100 kHz bandwidth; (iii) transimpedance of 2.25 GΩ; (iv) power consumption of 41 mW per channel including the ADC. The proposed architecture is one of the faster transimpedance amplifiers in the literature and it offers state-of-the-art noise performance.
A full design rationale, together with noise analysis, was presented describing all the trade-offs and design options. The system has two acquisition ranges (±200 pA and ±20 nA), while the acquisition bandwidths can be easily changed setting the OSR of the digital FIR filters. The front-end can apply a stimulus voltage to the BLM through virtual short-circuit of the first OTA. This feature is very useful when working with voltage-gated ion channels. Moreover, an integrated digital offset compensation loop balances any offsets in the Ag/AgCl electrodes.
The current-to-digital converter has been embedded in a multi-channel heterogeneous platform, along with microfluidic chips for current acquisitions from ion channels. Gramicidin-A, α-haemolysin and KcsA potassium channels have been used to prove both the platform and the current-to-digital converter.

Acknowledgments

This work was funded by the Engineering and Physical Sciences Research Council (EP/H044795/1). The authors would also like to acknowledge the generous gift of the KcsA protein from Phil Williamson and the help from Matteo Filippi. The data for this paper can be obtained from University of Southampton e-prints repository doi:10.5258/SOTON/394680.

Author Contributions

M.C., M.B. and M.T. designed the integrated CMOS transimpedance amplifier, performed electrical testing and noise measurements, and developed the noise model. S.C.S., M.d.P. and H.M. designed the microfluidic device and recorded data from ion channels. J.I. contributed the data acquisition from ion channels. All authors contributed to the writing and editing of the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hille, B. Ion Channels of Excitable Membranes, 3rd ed.; Sinauer Associates: Sunderland, MA, USA, 2001. [Google Scholar]
  2. Bayley, H.; Cremer, P.S. Stochastic sensors inspired by biology. Nature 2001, 413, 226–230. [Google Scholar] [CrossRef] [PubMed]
  3. Sakmann, B.; Neher, E. Single-Channel Recording, 2nd ed.; Springer-Verlag: Berlin, Germany, 2009. [Google Scholar]
  4. Fertig, N.; Blick, R.H.; Behrends, J.C. Whole cell patch clamp recording performed on a planar glass chip. Biophys. J. 2002, 82, 3056–3062. [Google Scholar] [CrossRef]
  5. Tien, H.T. Bilayer Lipid Membranes (BLM): Theory and Practice; Dekker: New York, NY, USA, 1974. [Google Scholar]
  6. Suzuki, H.; Takeuchi, S. Microtechnologies for membrane protein studies. Anal. Bioanal. Chem. 2008, 391, 2695–2702. [Google Scholar] [CrossRef] [PubMed]
  7. Kreir, M.; Farre, C.; Beckler, M.; George, M.; Fertig, N. Rapid screening of membrane protein activity: Electrophysiological analysis of OmpF reconstituted in proteoliposomes. Lab Chip 2008, 8, 587–595. [Google Scholar] [CrossRef] [PubMed]
  8. Suzuki, H.; Le Pioufle, B.; Takeuchi, S. Ninety-six-well planar lipid bilayer chip for ion channel recording fabricated by hybrid stereolithography. Biomed. Microdevices 2009, 11, 17–22. [Google Scholar] [CrossRef] [PubMed]
  9. Zagnoni, M. Miniaturised technologies for the development of artificial lipid bilayer systems. Lab Chip 2012, 12, 1026–1039. [Google Scholar] [CrossRef] [PubMed]
  10. Rosenstein, J. The promise of nanopore technology: Nanopore DNA sequencing represents a fundamental change in the way that genomic information is read, with potentially big savings. IEEE Pulse 2014, 5, 52–54. [Google Scholar] [CrossRef] [PubMed]
  11. Thei, F.; Rossi, M.; Bennati, M.; Crescentini, M.; Lodesani, F.; Morgan, H.; Tartagni, M. Parallel Recording of Single Ion Channels: A Heterogeneous System Approach. IEEE Trans. Nanotechnol. 2010, 9, 295–302. [Google Scholar] [CrossRef]
  12. Saha, S.C.; Thei, F.; de Planque, M.R.R.; Morgan, H. Scalable micro-cavity bilayer lipid membrane arrays for parallel ion channel recording. Sens. Actuators B Chem. 2014, 199, 76–82. [Google Scholar] [CrossRef]
  13. Crescentini, M.; Thei, F.; Bennati, M.; Saha, S.; de Planque, M.R.R.; Morgan, H.; Tartagni, M. A Distributed Amplifier System for Bilayer Lipid Membrane (BLM) Arrays With Noise and Individual Offset Cancellation. IEEE Trans. Biomed. Circuits Syst. 2015, 9, 334–344. [Google Scholar] [CrossRef] [PubMed]
  14. Crescentini, M.; Bennati, M.; Carminati, M.; Tartagni, M. Noise Limits of CMOS Current Interfaces for Biosensors: A Review. IEEE Trans. Biomed. Circuits Syst. 2013, 8, 278–292. [Google Scholar] [CrossRef] [PubMed]
  15. Kim, D.; Goldstein, B.; Tang, W.; Sigworth, F.J.; Culurciello, E. Noise Analysis and Performance Comparison of Low Current Measurement Systems for Biomedical Applications. IEEE Trans. Biomed. Circuits Syst. 2013, 7, 52–62. [Google Scholar]
  16. LeMasurier, M.; Heginbotham, L.; Miller, C. KcsA: It’s a potassium channel. J. Gen. Physiol. 2001, 118, 303–314. [Google Scholar] [CrossRef] [PubMed]
  17. Goldstein, B.; Kim, D.; Magoch, M.; Astier, Y.; Culurciello, E. CMOS low current measurement system for nanopore sensing applications. In Proceedings of the IEEE Biomedical Circuits and Systems Conference (BioCAS), San Diego, CA, USA, 10–12 November 2011; pp. 265–268.
  18. Axon Instruments, Inc. The Axon Guide for Electrophysiology & Biophysics Laboratory Techniques; Axon Instruments, Inc.: Foster City, CA, USA, 1993. [Google Scholar]
  19. Hsu, C.-L.; Venkatesh, A.G.; Jiang, H.; Hall, D.A. A hybrid semi-digital transimpedance amplifier for nanopore-based DNA sequencing. In Proceedings of the IEEE Biomedical Circuits and Systems Conference (BioCAS), Lausanne, Switzerland, 22–24 October 2014; pp. 452–455.
  20. Jafari, H.M.; Genov, R. Chopper-Stabilized Bidirectional Current Acquisition Circuits for Electrochemical Amperometric Biosensors. IEEE Trans. Circuits Syst. I Regul. Pap. 2013, 60, 1149–1157. [Google Scholar] [CrossRef]
  21. Rosenstein, J.; Wanunu, M.; Merchant, C.A.; Drndic, M.; Shepard, K.L. Integrated nanopore sensing platform with sub-microsecond temporal resolution. Nat. Methods 2012, 9, 487–492. [Google Scholar] [CrossRef] [PubMed]
  22. Schreier, R.; Temes, G.C. Understanding Delta-Sigma Data Converters; Wiley-IEEE Press: Piscataway, NJ, USA, 2004. [Google Scholar]
  23. Crescentini, M.; Rossi, M.; Bennati, M.; Thei, F.; Baschirotto, A.; Tartagni, M. A nanosensor interface based on delta-sigma Arrays. In Proceedings of the 3rd International Workshop on Advances in Sensors and Interfaces (IWASI), Trani, Italy, 25–26 June 2009; pp. 163–167.
  24. Shinwari, M.W.; Zhitomirsky, D.; Deen, I.A.; Selvaganapathy, P.R.; Deen, M.J.; Landheer, D. Microfabricated reference electrodes and their biosensing applications. Sensors 2010, 10, 1679–1715. [Google Scholar] [CrossRef] [PubMed]
  25. Matsumoto, T.; Ohashi, A.; Ito, N. Development of a micro-planar Ag/AgCl quasi-reference electrode with long-term stability for an amperometric glucose sensor. Anal. Chim. Acta 2002, 462, 253–259. [Google Scholar] [CrossRef]
  26. Polk, B.J.; Stelzenmuller, A.; Mijares, G.; MacCrehan, W.; Gaitan, M. Ag/AgCl microelectrodes with improved stability for microfluidics. Sens. Actuators B Chem. 2006, 114, 239–247. [Google Scholar] [CrossRef]
  27. Thei, F.; Rossi, M.; Bennati, M.; Crescentini, M.; Tartagni, M. An automatic offset correction platform for high-throughput ion-channel electrophysiology. Procedia Eng. 2010, 5, 816–819. [Google Scholar] [CrossRef]
  28. Enz, C.; Temes, G. Circuit techniques for reducing the effects of op-amp imperfections: Autozeroing, correlated double sampling, and chopper stabilization. Proc. IEEE 1996, 84, 1584–1614. [Google Scholar] [CrossRef]
  29. Fischer, J.H. Noise sources and calculation techniques for switched capacitor filters. IEEE J. Solid-State Circuits 1982, 17, 742–752. [Google Scholar] [CrossRef]
  30. Phillips, J.; Kundert, K. An Introduction to Cyclostationary Noise: Noise in mixers, oscillators, samplers, and logic. In Proceedings of the IEEE 2000 Custom Integrated Circuits Conference (CICC), Orlando, FL, USA, 21 May 2000; pp. 431–438.
  31. Gardner, W.A. Introduction to Random Processes with Applications to Signal & Systems, 2nd ed.; McGraw-Hill: New York, NY, USA, 1990. [Google Scholar]
  32. Proakis, J.G. Digital Communications; McGraw-Hill: New York, NY, USA, 2000. [Google Scholar]
  33. Woolley, G.A.; Wallace, B.A. Model ion channels: Gramicidin and alamethicin. J. Membr. Biol. 1992, 129, 109–136. [Google Scholar] [PubMed]
  34. Allen, T.W.; Andersen, O.S.; Roux, B. Energetics of ion conduction through the gramicidin channel. Proc. Natl. Acad. Sci. USA 2004, 101, 117–122. [Google Scholar] [CrossRef] [PubMed]
  35. Hromada, L.P.; Nablo, J.P.; Kasianowicz, J.J.; Gaitan, M.A.; DeVoe, D.L. Single molecule measurements within individually membrane- bound ion channels using a polymer-based bilayer lipid membrane chip. Lab Chip 2008, 8, 602–608. [Google Scholar] [CrossRef] [PubMed]
  36. Kawano, R.; Osaki, T.; Sasak, H.; Takinoue, M.; Yoshizawa, S.; Takeuchi, S. Rapid detection of a cocaine-binding aptamer using biological nanopores on a chip. Am. Chem. Soc. J. 2011, 133, 8474–8477. [Google Scholar] [CrossRef] [PubMed]
  37. Ferrari, G.; Farina, M.; Guagliardo, F.; Carminati, M.; Sampietro, M. Ultra-low-noise CMOS current preamplifier from DC to 1 MHz. Electron. Lett. 2009, 45, 1278–1280. [Google Scholar] [CrossRef]
  38. Ferrari, G.; Gozzini, F.; Molari, A.; Sampietro, M. Transimpedance amplifier for high sensitivity current measurements on nanodevices. IEEE J. Solid-State Circuits 2009, 44, 1609–1616. [Google Scholar] [CrossRef]
  39. Goldstein, B.; Kim, D.; Rottigni, A.; Xu, J.; Vanderlick, T.K.; Culurciello, E. CMOS low current measurement system for biomedical applications. In Proceedings of the IEEE International Symposium of Circuits and Systems (ISCAS), Rio de Janeiro, Brazil, 15–18 May 2011; pp. 1017–1020.
  40. Dai, S.; Rosenstein, J.K. A dual-mode low-noise nanosensor front-end with 155-dB dynamic range. In Proceedings of the IEEE Biomedical Circuits and Systems Conference (BioCAS), Atlanta, GA, USA, 22–24 October 2015; pp. 1–4.
Figure 1. (A) Diagram showing the patch clamp technique where a glass pipette is used to pull a patch of cell membrane. A low-noise transimpedance amplifier measures channel currents; (B) the planar bilayer membrane (BLM) technique where a suspended lipid bilayer contains an ion-channel. Again the current is read by a low-noise transimpedance amplifier. The picture also shows the electrical equivalent model of the BLM, consisting of a high value resistor (of the order of GΩ or greater) in parallel with a capacitance CS.
Figure 1. (A) Diagram showing the patch clamp technique where a glass pipette is used to pull a patch of cell membrane. A low-noise transimpedance amplifier measures channel currents; (B) the planar bilayer membrane (BLM) technique where a suspended lipid bilayer contains an ion-channel. Again the current is read by a low-noise transimpedance amplifier. The picture also shows the electrical equivalent model of the BLM, consisting of a high value resistor (of the order of GΩ or greater) in parallel with a capacitance CS.
Sensors 16 00709 g001
Figure 2. (a) Photograph of the 12-channel parallel recording platform highlighting each element: three small PCBs with two CMOS current-to-digital amplifiers described in this paper, three 4-channel microfluidic devices [12], and a PCB with FPGA and USB interface which is housed in the metal box; (b) Photograph of the platform showing board connections; (c) Photograph of the final platform with the metal box used for shielding.
Figure 2. (a) Photograph of the 12-channel parallel recording platform highlighting each element: three small PCBs with two CMOS current-to-digital amplifiers described in this paper, three 4-channel microfluidic devices [12], and a PCB with FPGA and USB interface which is housed in the metal box; (b) Photograph of the platform showing board connections; (c) Photograph of the final platform with the metal box used for shielding.
Sensors 16 00709 g002
Figure 3. Block diagram of the system. A lipid bilayer is formed in a microfluidic chip, with integrated Ag/AgCl electrodes. Reference Electrode (RE) is tight to VCM while working electrode (WE) is connected to the input of the transimpedance amplifier. The CMOS transimpedance amplifier acquires the input current iIN and digitizes it into a 1-bit high-frequency delta-sigma modulated stream. It also compensates for electrode and opamp offset by means of a digital compensation loop that is activated at the beginning of every experiment. The ADC output is filtered and decimated by a digital FIR filter implemented on a FPGA. Data communication with PC is via a USB link. Virtual short circuit realized by the input LNA is used to apply a stimulus voltage vSTIM to the BLM.
Figure 3. Block diagram of the system. A lipid bilayer is formed in a microfluidic chip, with integrated Ag/AgCl electrodes. Reference Electrode (RE) is tight to VCM while working electrode (WE) is connected to the input of the transimpedance amplifier. The CMOS transimpedance amplifier acquires the input current iIN and digitizes it into a 1-bit high-frequency delta-sigma modulated stream. It also compensates for electrode and opamp offset by means of a digital compensation loop that is activated at the beginning of every experiment. The ADC output is filtered and decimated by a digital FIR filter implemented on a FPGA. Data communication with PC is via a USB link. Virtual short circuit realized by the input LNA is used to apply a stimulus voltage vSTIM to the BLM.
Sensors 16 00709 g003
Figure 4. (a) Diagram showing the microfluidic device. The counter electrode sets the potential of the top fluid chamber, which is common to every BLM. There are four separate microcavities each with individual bilayers and separate integrated Ag/AgCl electrodes; (b) A photograph of the microfluidic device.
Figure 4. (a) Diagram showing the microfluidic device. The counter electrode sets the potential of the top fluid chamber, which is common to every BLM. There are four separate microcavities each with individual bilayers and separate integrated Ag/AgCl electrodes; (b) A photograph of the microfluidic device.
Sensors 16 00709 g004
Figure 5. (a) Block scheme of the proposed frontend. The signal direct path is composed of integrator, voltage amplifier, differentiator and active LPF; (b) Full schematic diagram of the frontend, where a subtractor has been added to eliminate derivative component of stimulus signal from the final output. This subtractor stage can be activated/deactivated using the control signal Sub; (c) Timing behavior of signals F1, F2 and F3.
Figure 5. (a) Block scheme of the proposed frontend. The signal direct path is composed of integrator, voltage amplifier, differentiator and active LPF; (b) Full schematic diagram of the frontend, where a subtractor has been added to eliminate derivative component of stimulus signal from the final output. This subtractor stage can be activated/deactivated using the control signal Sub; (c) Timing behavior of signals F1, F2 and F3.
Sensors 16 00709 g005
Figure 6. Timing behavior of the system. The frontend resets every period TR = 102.4 μs. During the reset, the output voltage is held at a constant value. AD sampling frequency is n-times higher than the reset frequency.
Figure 6. Timing behavior of the system. The frontend resets every period TR = 102.4 μs. During the reset, the output voltage is held at a constant value. AD sampling frequency is n-times higher than the reset frequency.
Sensors 16 00709 g006
Figure 7. Schematic diagram of the second-order ΔΣ converter. VCM is half the power supply.
Figure 7. Schematic diagram of the second-order ΔΣ converter. VCM is half the power supply.
Sensors 16 00709 g007
Figure 8. Architecture of the offset compensation loop. The offset compensation voltage VOFF is generated comparing the output of the frontend vOUT with VCM and then incrementing a 8-bit counter. Stimulus signal vSTIM comes from the FPGA and is summed to VOFF and fed to a DAC so as to create voltage vC A passive LPF filter at DAC output limits the noise.
Figure 8. Architecture of the offset compensation loop. The offset compensation voltage VOFF is generated comparing the output of the frontend vOUT with VCM and then incrementing a 8-bit counter. Stimulus signal vSTIM comes from the FPGA and is summed to VOFF and fed to a DAC so as to create voltage vC A passive LPF filter at DAC output limits the noise.
Sensors 16 00709 g008
Figure 9. (a) Simplified block scheme of the proposed architecture. Noise analysis is based on this system simplification where sampling is decoupled from the Sallen-Key filter; (b) Qualitative sketch of the noise PSD at node x; (c) Qualitative sketch of noise autocorrelation function at node x.
Figure 9. (a) Simplified block scheme of the proposed architecture. Noise analysis is based on this system simplification where sampling is decoupled from the Sallen-Key filter; (b) Qualitative sketch of the noise PSD at node x; (c) Qualitative sketch of noise autocorrelation function at node x.
Sensors 16 00709 g009
Figure 10. Effect the reset period T on the input-referred r.m.s. noise. Comparison of the proposed mathematical model with SpectreRF® pss-noise analysis.
Figure 10. Effect the reset period T on the input-referred r.m.s. noise. Comparison of the proposed mathematical model with SpectreRF® pss-noise analysis.
Sensors 16 00709 g010
Figure 11. Chip microphotograph.
Figure 11. Chip microphotograph.
Sensors 16 00709 g011
Figure 12. (a) Open-input input-referred noise measured at two different ranges (20 nA and 200 pA). The noise is as low as 4 fA/√Hz in the best condition (i.e., range = 200 pA and digital filtering at 10 kHz with a sinc3 LPF), while raises to 6 fA/√Hz at 200 pA range and digital filtering at 200 kHz. The highest noise floor, that is 40 fA/√Hz, is shown in the wider 20 nA range. Digital FIR filters with triangular windowing and 150 taps were used; (b) Short-time trace recorded at bandwidth = 175 kHz showing periodic spikes every TR = 102.4 μs. This spikes are linked to the period reset activity; (c) Relation between input-referred r.m.s. noise at 10 kHz and input capacitance Cin, which takes into care microfluidic device together with BLM. Measured values are compared with simulations and theoretical estimation, validating Equation (19).
Figure 12. (a) Open-input input-referred noise measured at two different ranges (20 nA and 200 pA). The noise is as low as 4 fA/√Hz in the best condition (i.e., range = 200 pA and digital filtering at 10 kHz with a sinc3 LPF), while raises to 6 fA/√Hz at 200 pA range and digital filtering at 200 kHz. The highest noise floor, that is 40 fA/√Hz, is shown in the wider 20 nA range. Digital FIR filters with triangular windowing and 150 taps were used; (b) Short-time trace recorded at bandwidth = 175 kHz showing periodic spikes every TR = 102.4 μs. This spikes are linked to the period reset activity; (c) Relation between input-referred r.m.s. noise at 10 kHz and input capacitance Cin, which takes into care microfluidic device together with BLM. Measured values are compared with simulations and theoretical estimation, validating Equation (19).
Sensors 16 00709 g012
Figure 13. (a) Functionality of the OCL. An offset in the nA range usually exists at the beginning of the experiment. The OCL changes the DC component of the applied stimulus vC until a zero current is recorded (i.e., VOUT = VCM); (b) Functionality of the subtractor. A 10 pF capacitor is connected to the input while a 100 mVpp triangular wave at 200 Hz is applied as vC signal leading to a 800 pApp current square wave flowing through the input. The input current is overestimated when the subtractor is turned-off; (c) Effect of the OCL on the input-referred noise. When the OCL is activated and a stimulus signal vSTIM is applied, the noise rises from 4 fA/√Hz to 6 fA/√Hz due to noise sources in the OCL. Note that vC is filtered with an external 10 nF capacitance.
Figure 13. (a) Functionality of the OCL. An offset in the nA range usually exists at the beginning of the experiment. The OCL changes the DC component of the applied stimulus vC until a zero current is recorded (i.e., VOUT = VCM); (b) Functionality of the subtractor. A 10 pF capacitor is connected to the input while a 100 mVpp triangular wave at 200 Hz is applied as vC signal leading to a 800 pApp current square wave flowing through the input. The input current is overestimated when the subtractor is turned-off; (c) Effect of the OCL on the input-referred noise. When the OCL is activated and a stimulus signal vSTIM is applied, the noise rises from 4 fA/√Hz to 6 fA/√Hz due to noise sources in the OCL. Note that vC is filtered with an external 10 nF capacitance.
Sensors 16 00709 g013
Figure 14. (a) Data from gramicidin-A channels. Applied potential = 100 mV and the acquisition has been performed at 625 Hz; (b) Current trace showing the insertion of two α-HL nanopores. Experiments performed with 2.5 μg/ml of α-HL protein in 1 M of KCl solution. Acquisition performed at 200 pA range with post-processing filtering at 1.25 kHz; (c) KcsA recording at 125 mV with data filtered at 5 kHz and (d) 10 kHz. The buffer was 150 mM KCl, pH 4.0.
Figure 14. (a) Data from gramicidin-A channels. Applied potential = 100 mV and the acquisition has been performed at 625 Hz; (b) Current trace showing the insertion of two α-HL nanopores. Experiments performed with 2.5 μg/ml of α-HL protein in 1 M of KCl solution. Acquisition performed at 200 pA range with post-processing filtering at 1.25 kHz; (c) KcsA recording at 125 mV with data filtered at 5 kHz and (d) 10 kHz. The buffer was 150 mM KCl, pH 4.0.
Sensors 16 00709 g014
Figure 15. Input-referred noise compared with benchmark instrument for biological current acquisition. The peaks shown in all the lines are due to spurious coupling from the 50 Hz powerline.
Figure 15. Input-referred noise compared with benchmark instrument for biological current acquisition. The peaks shown in all the lines are due to spurious coupling from the 50 Hz powerline.
Sensors 16 00709 g015
Table 1. Values of system parameters.
Table 1. Values of system parameters.
ComponentC1C2C3C4R4R3TRτR
Value1 pF22 pF1 pF102.4 pF1 MΩ2.2 kΩ102.4 μs4.8 μs
Table 2. Comparison of noise theory with simulation and measures.
Table 2. Comparison of noise theory with simulation and measures.
Theory (18)SimulationMeasure
RMS NOISE at 10 kHz380 fA400 fA420 fA
ConditionsCIN = CP = 3 pFCIN = CP = 3 pFOpen-input
Table 3. State-of-the-art comparison.
Table 3. State-of-the-art comparison.
PaperNoise floor @ Room TemperatureEmbedded ADCAnalog Power ConsumptionDigital Power ConsumptionInput Capacitance for Characterization of Noise FloorOperating Bandwidth [kHz]Gain [GΩ]Technology Node
[21]12 fA/√HzNO---10.000<1CMOS 0.13 μm
[20]2 fA/√HzNO3 μW--6>1CMOS 0.13 μm
[37]0.5 fA/√HzNO--1 pF1000*1-
[39]6 fA/√HzNO1.5 mW-7 pF50-CMOS 0.5 μm
[38]4 fA/√HzNO45 mW-800 fF10,0000.06CMOS 0.35 μm
[40]11.6 fA/√HzNO5.22 mW--14000.01CMOS 0.18 μm
[13]3 fA/√Hz@B = 625 Hz
12 fA/√Hz @B = 10 kHz
YES20 mW20 mW3 pF102.25CMOS 0.35 μm
This work4 fA/√Hz*2 @B = 7.5 kHz
6 fA/√Hz *2 @B = 175 kHz
YES21 mW20 mW3 pF1002.25CMOS 0.35 μm
*1 The output of this amplifier is a current; *2 Noise floor measured at ±200 pA range.

Share and Cite

MDPI and ACS Style

Crescentini, M.; Bennati, M.; Saha, S.C.; Ivica, J.; De Planque, M.; Morgan, H.; Tartagni, M. A Low-Noise Transimpedance Amplifier for BLM-Based Ion Channel Recording. Sensors 2016, 16, 709. https://doi.org/10.3390/s16050709

AMA Style

Crescentini M, Bennati M, Saha SC, Ivica J, De Planque M, Morgan H, Tartagni M. A Low-Noise Transimpedance Amplifier for BLM-Based Ion Channel Recording. Sensors. 2016; 16(5):709. https://doi.org/10.3390/s16050709

Chicago/Turabian Style

Crescentini, Marco, Marco Bennati, Shimul Chandra Saha, Josip Ivica, Maurits De Planque, Hywel Morgan, and Marco Tartagni. 2016. "A Low-Noise Transimpedance Amplifier for BLM-Based Ion Channel Recording" Sensors 16, no. 5: 709. https://doi.org/10.3390/s16050709

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop