Вы находитесь на странице: 1из 71

Simisola O.

Oludare

Bioe 431: Instrumentation Lab I

Section 14163

TA: Yanjun Chen

Introduction to Electronic Equipment

Abstract

The purpose of this lab is to gain an understanding of how a function generator and an
oscilloscope work and then use that knowledge to perform a series of experiments. The
experiments which will be performed in this laboratory are to determine the value of an unknown
resistor, determine the input impedance of the function generator and the output impedance of
the oscilloscope, and to quantify the function of a high pass and low pass filter. The unknown
resistance was determined to be 84Ω by building a circuit in which the unknown resistor was in
series with a known resistor. The potential difference across the circuit was then measured and
used to calculate the value of the resistor. The input and output impedance of the function
generator and the oscilloscope were determined in a similar manner. The function generator and
the oscilloscope were considered elements built in series with a known resistor. The potential
difference across the circuit was then measured and used to calculate the value of the output and
input impedance. The output impedance was 51.34 Ω and the input was 19,680 Ω. As expected,
the input impedance was orders of magnitude greater than the output impedance. Finally, the
frequency response, cutoff frequency and step response of a low pass filter (100 kΩ and 10 nF)
and a high pass filter (100 kΩ and 10 nF) were determined. The cutoff frequency for the low pass
was 1.8 kHz and the cutoff frequency for the high pass was 171 Hz. As expected, the low pass
filter attenuated high frequency signals and amplified low frequency signals. Conversely, the
high pass filter attenuated low frequency signals and amplified high frequency signals. Overall,
the goals of this lab were met and each experiment produced expected results. Through these
series of experiments, we were able to gain an understanding of how simple circuits work and an
appreciation for how proper measurements and mathematical models can be used to determine
the function of a circuit.

Introduction

An electronic circuit is an instrument composed of many electronic elements arranged in a


particular configuration to perform specific, sometimes complex, functions such as signal
amplification and signal processing. To understand the function of an electronic circuit, it needs
to be provided an input current while simultaneously measuring its output current (or voltage).
The source of the current can either be a battery providing direct current (DC), an alternating
current (AC) from a power socket, or a function generator providing either DC or AC. The

1
output can then be measured either by an ammeter which measures current or an oscilloscope
which measures potential difference across the circuit. For the series of experiments performed
in this laboratory, we will be using a function generator and an oscilloscope to provide input and
output, respectively. The function generator outputs a variety of voltage signals over a range of
frequencies. The waveforms which will be used in these experiments will be sine and square waves
but the function generator is also capable of producing triangular, sawtooth, pulse, noise, and
arbitrary shape signals which can either be bursts of the signal (single pulse), or a periodic signal.
The oscilloscope is measurement equipment with a preamplifier capable of measuring and displaying
the AC, DC and ground output signals.

Before using the function generator and oscilloscope to study a circuit, it is important to note that
neither are ideal conductors. In the first experiment, we will find out the inherent resistances of
both the function generator and the oscilloscope. The resistance of the function generator is the
output impedance and it is desirable that this resistance is considerably low in order to minimize
the effort needed to push the current through the function generator. Conversely, the input
impedance is the resistance of the oscilloscope and it is desirable that this resistance is orders of
magnitude higher than the output impedance. To determine the input and the output impedances
of the function generator and the oscilloscope, respectively, we will design a couple of circuits
with resistors in series to the function generator and the oscilloscope. The potential difference
across the circuits will be measured and the equations below will be used to determine the input
and output impedances.

Figure 1. Variation to Ohm's law which allows us to calculate the input and output impedances from input and output
voltages.
Vi is the input voltage, Vx is the output voltage, RO is the output impedance and Zi is the input impedance.

In the second experiment of the lab, we will be using these tools to determine the value of an
unknown resistor. To determine the value of the unknown resistor, we use a variation of Ohm’s
law which allows us to calculate the resistance of the circuit after measuring the input and output
voltages of the circuit (Fig 2). This variation to Ohm’s law is necessary because without an
ammeter measuring the current across the circuit it is impossible to use the traditional
formulation. Using the oscilloscope, we are able to measure the input and output voltages across
the circuit in pseudo-differential mode, a configuration which allows the voltage across the
circuit to be measured in relation to the positive and negative leads of the oscilloscope.

2
Figure 2. Variation to Ohm's law which allows us to calculate the unknown resistance from input and output voltages.
Vi is the input voltage, Vx is the output voltage, R is the known resistor and RU is the unknown resistance.

In the final experiment, we will be studying the functions and properties of a passive low pass
and a passive high pass filter. Both of these filters are first order Resistance-Capacitance (RC)
circuits which are composed of a resistor and a capacitor in series. The function of a low pass
filter is to attenuate high frequency signals and amplify low frequency signals. This behavior
occurs in a low pass filter because the current from the source flows through the capacitor first,
where the voltage is differentiated then directed through the resistor. Conversely, the high pass
filter acts an integrator where the current which flows from the source through the resistor then
capacitor is integrated. The function of a high pass filter is to attenuate low frequency signals and
amplify high frequency signals. Thus, it is expected that these filters will behavior differently to
a range of frequencies. To understand the behavior of these filters, we will determine their
frequency response, cutoff frequency and step response. The frequency response of a circuit is
the response of a circuit to a sinusoid input. To create a map of the circuit’s frequency response,
input signals of various frequencies are run through the circuit and expressed in terms of gain
(output voltage/input voltage) in decibels (dB) vs. frequency and phase shift ( input to output
time delay expressed in degrees) vs. frequency. Through this mapping, the frequency at which
the input signal is being attenuated can be determined. This is known as the cutoff frequency.
Finally, to characterize the filters, we will determine the step response at their respective cutoff
frequencies allowing us to determine if the filter is attenuating or amplifying the input and the
respective time constant.

Through these series of experiments, we will gain an understanding of how simple circuits work
and an appreciation for how proper measurements and mathematical models can be used to
determine the function of a circuit.

Methods

Equipment Familiarization

To accomplish the goals of this lab, we had to understand the controls and display panel of the
oscilloscope and the function generator. The oscilloscope and function generator were studied by
generating a sine wave of 2 V, DC offset of 0 V and frequency of 500 Hz on the function
generator. The function generator was then connected to the oscilloscope to view the generated

3
sine wave. The voltage range, frequency range and DC voltage range of the function generator
were determined. The range of sweep speeds and the range of voltages which can be viewed on
the oscilloscope were also recorded. Additionally, the error from the function generator was also
determined by generating a 10 V sine wave and measuring the output on the oscilloscope.

Determining the current and resistance of a circuit

The resistance of an unknown resistor was determined by building a circuit with the unknown
resistor in series with a 56 kΏ resistor. A 0.07 mA alternating current (10.2V) was then sent
from the function generator through the circuit. The voltage drop across the circuit was measured
by the oscilloscope in pseudo-differential mode (Fig 3), allowing us to measure the potential
difference between the positive end and the negative input leads regardless of their relation to
ground. This step was necessary because all of the oscilloscopes in the laboratory are non-
differential and measure potential difference relative to ground. After measuring the potential
difference across the circuit, the equation above used to calculate the value of the unknown
resistor (Fig. 2).

Input and Output Impedance

The input impedance and the output impedance of the function generator and oscilloscope,
respectively, were determined by building a circuit in which both the function generator and
oscilloscope were in series with a known resistor of 820 kΩ and 30 Ω, respectively (Fig 4). The
input impedance was assumed to be purely resistive. An alternating current from the function
generator at 10.0 V was sent through both circuits and the resulting potential difference was
measured. The input and output impedances were then calculated using the equations from
section 3 in the pre lab.

Low Pass Filter

The passive low pass filter was built using a resistor of 100 kΩ and a capacitor of 10 nF arranged
in series. The capacitor followed the resistor and served as the output terminal. The cutoff
frequency of this circuit was determined by delivering alternating currents (10.0V) from the
function generator at frequencies spaced at logarithmic intervals ranging from 60 Hz – 50,000
kHz (Fig 6). The step response of the circuit was then determined by delivering a square wave
equal to the cutoff frequency through the circuit.

High pass filter

The passive high pass filter was built using a resistor of 100 kΩ and a capacitor of 10 nF
arranged in series. The resistor followed the capacitor and served as the output terminal. The
cutoff frequency of this circuit was determined by delivering alternating currents (10.0V) from
the function generator at frequencies spaced at logarithmic intervals ranging from 10 Hz –

4
50,000 kHz (Fig 8). The step response of the circuit was then determined by delivering a square
wave equal to the cutoff frequency through the circuit.

Results

Equipment Familiarization

1. The range for peak to peak voltage (Vpp) available for a sine wave of a function generator is 1
– 20 V. This range is consistent across all frequencies.
2. The frequency can be varied over 1 mHz – 10.000,000 MHz.
3. The range for DC voltage is -9.50 V – 9.50 V. The high level must be greater than the lo
level.
4. The range of the sweep speed, from fastest to slowest, is 5.000 ns/div – 50 s/div.
5. The range of voltages which can be displayed on the oscilloscope screen is 20.00 mV/div –
100.00 V/div.
6. The output signal of the function generator is 10.2 Vpp, resulting in an error of 0.2 Vpp.
7. At 100 µs/div, 1 cycle is visible on the oscilloscope. At 10 ms/div, 60 cycles are visible on
the oscilloscope.

Determining the current and resistance of a circuit

Figure 3. Psuedo-differential mode to measure unknown resistance.

1. The unknown resistor had a resistance of 84 kΩ.


2. The resistance of the unknown resistor was calculated using a known resistor of 56 kΩ. If the
resistance of the known resistor was much greater than the unknown resistance, the known
resistor would have acted as an open circuit and current would not have passed through the
unknown resistor. If the resistance of the known resistor was much lower than the known
resistance, the known resistor would have caused a short circuit.

5
Input Impedance of the Function Generator and Output Impedance of the Oscilloscope

Figure 4. Measuring output impedance of the oscilloscope

1. The output impedance of the function generator was determined to be 51.34 Ω.

Figure 5. Measuring input impedance of the function generator

2. The input impedance of the oscilloscope was determined to be 19,680 kΩ.

6
Low Pass Filter

1. The corner frequency of the filter is approximately 1.8 kHz.

-10
Gain (dB)

-20

-30

-40
1 10 100 1000 10000 100000
Frequency (Hz)

0
Phase (degrees)

-20
-40
-60
-80
-100
1 10 100 1000 10000 100000
Frequency (Hz)

Figure 6. Low pass filter Bode Plot (a) Gain vs. Frequency (b) Phase vs. Frequency.

120 10
100
5
Voltage (5V)
Voltage (V)

80
60 0
40
-5
20
0 -10
0 200 400 600 0 200 400 600
Time (100.0 µs/) Time (100.0 µs/)

Figure 7. Low pass filter step response to a square wave at the cutoff frequency (a) Step
Input. (b) Step Response.

7
High Pass Filter

1. The corner frequency of the filter is approximately 171 Hz.

0
-2
-4
Gain (dB)

-6
-8
-10
-12
1 10 100 1000 10000 100000
Frequency (Hz)

80
60
Phase (degrees)

40
20
0
-20
-40
1 10 100 1000 10000 100000
Frequency (Hz)

Figure 8. High pass filter Bode Plot (a) Gain vs. Frequency (b) Phase vs. Frequency.

120 30
100 20
Voltage (V)

Voltage (V)

80 10
60 0
40 -10
20 -20
0 -30
0 2 4 6 0 2 4 6
Time (2.000 ms/) Time (2.000 ms/)

Figure 9. High pass filter step response to a square wave at the cutoff frequency (a) Step
Input. (b) Step Response. (a) Step Input. (b) Step Response.

8
Discussion

The purpose of this lab was to use a function generator and oscilloscope to understand the
function of various circuits and circuit elements. In the first experiment, we studied the input and
output impedances of the oscilloscope and function generator, respectively. The ratio of input to
output impedance was determined to be 383.3 k. This ratio agrees with the expectation that the
input impedance will be orders of magnitude greater than the output impedance. The great
difference in input and output impedance allows current to be sent through the circuit efficiently
and the oscilloscope to measure the output voltage without affecting the signal.

The bandwidth measured for the low pass filter was approximately 1800 kHz. This means that at
lower frequencies, the gain is close to 0 resulting in a preservation of the signal and the phase is
greater than -45˚ resulting in minimal time delay of the signal. However, at higher frequencies,
the gain is less than 0 resulting in the attenuation of the signal and the phase is less than -45˚
resulting in greater time delay of the signal. The step response of the low pass filter to the
measured cutoff frequency resulted in a charging step response. These results are due to the
orientation of the capacitor and resistor. By placing the capacitor before the resistor, the circuit is
discharging the circuit at higher frequencies and charging it at lower frequencies.

As expected the high pass filter behaved inversely to the low pass filter. The bandwidth
measured for the high pass filter was 161 Hz, meaning that at higher frequencies the gain is close
to 0 resulting in an a preservation of the signal of and the phase is less than 45˚ resulting in
minimal time delay of the signal. But at lower frequencies, the gain is less than 0 resulting in the
attenuation of the signal and the phase is greater than 45˚ resulting in a time greater time delay.
These results are due to the orientation of the capacitor and resistor. Given the resistance and
capacitance of the RC circuit, we were able to calculate the expected cutoff frequency of the high
pass filter. The calculated cutoff frequency was 159.2 Hz, resulting in an error of 5.8%.
However, this does not drastically change the step response of the high pass filter. The step
response of the high pass filter to the measured cutoff frequency resulted in a discharging step
response. By placing the resistor before the capacitor, the circuit is discharging the circuit at
lower frequencies and charging it at higher frequencies.

Together, the series of experiments in this lab allowed us to gain an understanding and
appreciation for measuring electronic signals from different circuits and circuit elements.
Additionally, we were able to use mathematical models such as Ohm’s law and the frequency
response of an RC circuit to verify the results.

9
Simisola O. Oludare

Bioe 431: Instrumentation Lab I

Section 14163

TA: Yanjun Chen

Digital Data Acquisition Using Lab View

Abstract

The purpose of this lab is to explore the process of analog data collection, analog to digital
conversion and digital analysis of analog signals. The instruments used were the USB-6009 Data
Acquisition Module hardware and LabVIEW software. The first experiments in the series were
used to explore sampling and the Nyquist criterion by sampling a few analog signals and
observing the results. In the second experiment, we explored the FFT and plot interpolation. And
in the last experiment, we synthesized a 100 Hz (50% duty cycle) square wave by summing
sinusoids of the first two, three and four harmonics of 100 Hz. Through these experiments, we
learned that Nyquist is often insufficient and that aliasing occurs when a signal is either under
sampled or oversampled. Additionally, we were able to generate a square wave by analyzing its
frequency content. Overall, this lab showed us the importance of sampling at an appropriate
frequency given the dynamic range of the signal and also the power of the Fourier Transform to
generate and recreate signals.

Key words

Analog signals, Sampling, Analog to Digital Conversion, Nyquist, Fast Fourier Transform,
Aliasing

Introduction

By virtue of occurring in the non-digital world, biological signals such as heart rate and muscle
activity are analog meaning that they are continuous and can achieve any value within their
output range (i.e. dynamic range). However, due to the development of digital computational
tools which are faster, more controllable and easier to use compared to their analog counterparts,
digital processing and analysis of analog signals has become the norm. Although digital methods
are the standard way to analyze and process analog signals, analog sensors still remain the gold
standard for collecting analog signals.

To combine these two methods, it is necessary to convert the analog signals to digital signals
through a processed called sampling. Sampling of analog signals is performed by an analog-to-
digital (A-to-D) converter which collects the analog signal at a predetermined rate. The choice of
this rate is highly important and can determine whether or not the analog signal is digitized
accurately. The sampling rate is determined via Nyquist’s criterion which states that the
1
sampling rate of the A-to-D converter must be a least twice the frequency of the measured signal
in order for the signal to be digitized accurately and not undergo aliasing, thereby digitizing the
analog signal at a different frequency. Once the analog signal has been digitized correctly, it can
then be analyzed using a variety of mathematical and computer science techniques. One of those
techniques is the Fast Fourier Transform (FFT). The FFT is a discrete (i.e., digital) method of
determining the frequency components of a signal. This method is useful for developing filters
based on frequency content and mathematically synthesizing (or recreating) analog signals.

In this lab, we will be exploring the whole process with a series of experiments. The experiments
begin with the acquisition and sampling of analog signals, frequency analysis of digital signals
and synthesis of digital signals based on FFT.

Methods

The analog signals analyzed in these experiments were generated by an arbitrary waveform
generator and monitored via an oscilloscope. The analog signals were sampled using the USB-
6009 Data Acquisition Module hardware (National Instruments, Austin, TX) and LabVIEW
software (National Instruments, Austin, TX). The USB-6009 data acquisition unit has a
maximum sampling rate of 48,000 Hz and a 14-bit A-to-D resolution. The Measurement and
Automation Explorer program within LabVIEW was used to operate the data acquisition unit.
The data acquisition unit was configured to continuous for acquisition mode and RSE for
terminal configuration, the ideal method for acquiring signals from floating battery sources. In
addition to enabling signal acquisition, LabVIEW is also useful for performing computation (via
a program) on acquired signals and also for generating digital signals. A program in LabVIEW is
called a virtual instrument (VI) and allows the user to view signals and perform computational
processes in real time.

In the first experiment of the lab, we explored sampling and the Nyquist criterion by sampling a
few analog signals and observing the results. First, we developed a 300 Hz signal (10 V p-p) and
sampled it at 300 Hz. After this, we sampled a 290 Hz signal at 300 Hz. Then, we sampled 150
Hz, 140 Hz and 160 Hz signals at 300 Hz. Finally, we sampled a 300 Hz at this series of
frequencies, fsampling = fsignal*n + 3 where n is equal to 1,2,3,4, respectively.

In the second experiment, we explored the FFT and plot interpolation. First, we sampled a 5 Hz
and 25 Hz at 5 times their frequency and then determined their frequency content using FFT.
Next, we generated a 100 Hz sine wave, a 100 Hz (50% duty cycle) square wave and a 100 Hz
(10% and 100% symmetry) triangle wave. While generating these signals, we performed a FFT
on them and recorded their power and phase spectra. Additionally, we generated a 200 Hz signal
and sampled it a 100 Hz, 200 Hz, 400 Hz and 800 Hz and observed the effects of linear and
square interpolation methods.

2
In the last experiment, we synthesized a 100 Hz (50% duty cycle) square wave by summing
sinusoids of the first two, three and four harmonics of 100 Hz. We did this by developing a
LabVIEW VI which generated the necessary number of sinusoids and summed them in real time.

Results

By sampling a 300 Hz sine wave at 300 Hz, we were only able to observe a fraction of a period
of the sine wave. This was due to sampling at a rate substantially below the Nyquist sampling
rate which is twice the signal being produced. When a 290 Hz sine wave was sampled at 300 Hz,
we observed a signal with a frequency of 10 Hz, a value equal to the frequency of the produced
signal (290 Hz) subtracted from the sampling frequency (300 Hz). When two sine waves of 140
Hz and 160 Hz, respectively, were sampled at 300 Hz, we observed a signal similar to a beat
signal where the amplitudes of both sampled signals (140 Hz and 160 Hz) were being modulated
by a sine wave of 10 Hz. By sampling a 300 Hz sine wave at the following sampling rates 303
Hz, 603 Hz, 903 Hz and 1203 Hz, the observed frequencies were 3.3 Hz, 307.7 Hz, 333 Hz and
307. 7 Hz, respectively. The mathematical relationship between the sampling frequencies
(fsampling), the observed frequency (fobs) and the input frequency (fsignal) is fsampling = fobs*n + fsignal
where n is equal to 1,2,3,4, respectively.

Figure 1. 5 Hz Sine Wave. (Left) Five seconds (i.e. 250 samples) (Right) Five cycles

3
Figure 2. 25 Hz Sine Wave. (Left) Five seconds (i.e. 625 samples) (Right) Five cycles

Figure 3. (Left) Top: Amplitude vs. Frequency plot of the 5 Hz sine wave, Bottom: Phase vs. Frequency plot of the 5 Hz
sine wave. (Right) Top: Amplitude vs. Frequency plot of the 25 Hz sine wave, Bottom: Phase vs. Frequency plot of the 25
Hz sine wave.

4
Figure 4A. (Top) Left: Linear interpolation of a 200 Hz sine wave sampled at 100 Hz, Right: Square interpolation of a 200
Hz sine wave sampled at 100 Hz. (Bottom) Left: Linear interpolation of a 200 Hz sine wave sampled at 200 Hz, Right:
Square interpolation of a 200 Hz sine wave sampled at 200 Hz.

Figure 4B. (Top) Left: Linear interpolation of a 200 Hz sine wave sampled at 400 Hz, Right: Square interpolation of a 200
Hz sine wave sampled at 200 Hz. (Bottom): Linear interpolation of a 200 Hz sine wave sampled at 800 Hz

5
Figure 5. (Left) Top: Amplitude vs. Frequency plot of the 100 Hz sine wave, Middle: Amplitude vs. Frequency plot of the
100 Hz square wave, Bottom: Amplitude vs. Frequency plot of the 100 Hz 10% asymmetric triangle wave. (Right) Top:
Phase vs. Frequency plot of the 100 Hz sine wave, Middle: Phase vs. Frequency plot of the 100 Hz square wave, Bottom:
Phase vs. Frequency plot of the 100 Hz 10% asymmetric triangle wave.

6
Figure 6. 100 Hz Square wave synthesized from sum of sinusoids (Top) First two harmonics (Middle) First three
harmonics (Bottom) First Four harmonics

Discussion

In this lab, we explored the processes of analog data collection, analog to digital conversion,
frequency analysis of acquired analog signals and synthesis of signals by summing sinusoids. By
sampling signals at rates below the Nyquist criterion, we were able to observe aliasing and

7
amplitude modulation. Additionally, sampling at a rate slightly greater than Nyquist (in this case,
20 Hz above Nyquist) also resulted in aliasing and amplitude modulation. We also observed that
the accuracy of the digitized signal depends both on the sampling rate and the form of
interpolation performed on the signal. The relationship between the observed frequency and the
sampling rate was determined to be fobs= (fsampling - fsignal)/n and we observed that the linear
interpolation method provided the best representation of the signal. Additionally, we were able to
synthesize a good approximation of a 100 Hz square wave after determining its frequency
content. However, we only used up to the first four harmonics and did not modify the
contribution of each sinusoid by using the Fourier series formula, limiting the accuracy of our
synthesized square wave.

In order to perform these series of experiments, particularly the FFT using LabVIEW, it was
necessary for us to convert the analog signals we generated to digital signals. This process
required us to gain a familiarization with signal acquisition and digitization. We discovered that
Nyquist (twice the sampled frequency) is often insufficient and that aliasing occurs when a signal
is either under sampled or oversampled. Aliasing resulted when we sampled both a 140 Hz and
160 Hz signal at 300 Hz. But when we sampled 5 Hz and 25 Hz signals at five times their
sampling frequencies (three orders of magnitude greater than Nyquist), the signals were
reproduced without aliasing and amplitude modulation. To us, this means that for sampling
biological signals such as human gait and heart rate, the sampling rate should be greater than
Nyquist in order to properly reproduce the results. However, this can result in the acquisition of
high frequency noise which can be observed using the FFT and filtered using a low pass or band
pass filter. By combining the FFT with the Fourier series, it is also possible to generate analog
signals such as the approximation of a square wave created in this lab. For biological signals, this
can be used to understand how particular frequencies affect the signal content by systematically
combining different frequencies in order to recreate the original signal. However, as observed in
this lab, the power of each sinusoid must be adjusted based on the power of the frequency on the
spectrum.

Through this lab, we learned the basics of analog data collection and digital manipulation of
analog signals. Although the signals generated in this lab were from an arbitrary waveform
generator, the methods learned in this lab can (with some difficulty) be applied to biological
signals.

8
Simisola O. Oludare

Bioe 431: Instrumentation Lab I

Lab partner: Nadia Crawley

Section 14163

TA: Yanjun Chen

Signal Processing Using the 741 OP-AMP and Digital Filters

Abstract

The purpose of this lab was to perform analog and digital signal processing using active analog
circuits and Butterworth filters, respectively. The active analog filters were created using a 741-
opertational amplifier and the Butterworth filter was created in LabVIEW. The lab consisted of
five experiments. In the first experiment of the series, we determined the frequency response of a
unity gain amplifier. In the second experiment, we determined the gain and bandwidth of two
non-unity gain amplifiers with theoretical gains of 20 and 100, respectively. In the third
experiment, we used a potentiometer to add DC offset to a 100 mV, 5 kHz signal. In the fourth
experiment, we designed and built a low pass, high pass and band pass filter with theoretical cut
off frequencies of 15 kHz, 1.59 kHz, and 15 kHz and 1.59 kHz, respectively. And then for each
filter, we determined the frequency response, the -3dB frequency and the step response at the -
3dB frequency. Based on our results, we discovered that analog circuits are not very accurate or
precise. The measured gain of the unity gain amplifier was less than one and decreased as the
frequency was increased. The range of DC offset was 40.34 V – 46.88 V. The measured cutoff
frequencies were not equal to the ideal cutoff frequencies. The measured cutoff frequencies were
13.1 kHz, 1.6 kHz and 5 kHz and 20 kHz for the low pass, high pass and band pass filters,
respectively. Unlike the analog filters, the digital filter was more precise. The cutoff frequency to
eliminate 60 Hz noise from a 30 Hz signal was 35 Hz and the signal to noise ratio was 31.6.
Through this lab, we were able to design and create various active analog circuits and compare
their real behavior to the ideal behavior. Additionally, we will be able to compare digital and
analog filtering methods.

Key words

Operational Amplifiers, Unity Gain, Non-Unity Gain, Analog Filters, Digital Filters, Butterworth
Filter, Convolution, High Pass Filter, Low Pass Filter, Band Pass Filter, Signal to Noise ratio

Introduction

Biologically generated electrical signals have very low amplitudes and are typically noisy due to
interfering biological and environmental signals. Thus, before it is analyzed, biological signals
need to be amplified and filtered. Together, the amplification and filtering processes are called
1
signal conditioning. These processes can be performed using analog circuits and digital
programming. By using mathematical tools based on Kirchhoff’s laws, it is possible to design
analog circuits to meet particular shifting, amplification and filtering specifications. And using
the Fourier Transform and convolution, it is possible to design digital filters to shape the
frequency content of a signal. In this lab, we will be designing and creating active analog circuits
to perform unity and non-unity gain amplification as well as filtering of high and low frequency
content. Additionally, we will also be designing multiple low pass digital filters.

The active analog circuits created in this lab will be done with the 741-operation amplifier (op-
amp) and a potentiometer. The 741 op-amp is an integrated circuit (IC) with high gain, high
input impedance and low output impedance. The 741 op-amp needs to be powered by 15 V (DC)
and grounded. Also, the 741 op-amp has both an inverting and non-inverting input for creating
inverting and non-inverting amplifiers, respectively. The inverting amplifier provides a gain that
is negative resulting in an output signal that is out of phase (time delayed). By virtue of its
design, the 741 op-amp has a very constrained bandwidth compared to passive amplifiers which
have an infinite bandwidth. However, the advantage of the limited bandwidth is that circuits
made with the op-amp are generally easier to tune. Another advantage of using the 741 op-amp
is that it can be easily combined with additional circuit components. For example, by combining
the 741-op amp with a potentiometer, the voltage of the signal can be increased by adding DC
offset (shifting) and by placing resistors in series, the signal can be amplified using the non-
inverting or inverting port of the op-amp. Additionally, by combining the 741 op-amp with
resistors and circuits (RC circuit), a low pass, high pass or band pass filter can be created. In this
lab, we will be creating all three types of filters.

An ideal low pass filter is an RC circuit which attenuates frequency signals above its cut off
frequency while an ideal high pass filter is an RC circuit which attenuates frequency signals
below its cut off frequency. And a band pass filter is a combination of both. Ideally, the
amplitude of the frequencies which are passed should have the same level of amplification.

In addition to the analog circuits, digital filters will be created in LabVIEW. The filters which
will be created in this lab will be 3rd and 5th order low pass Butterworth filters. The Butterworth
filter is a finite impulse response (FIR) filter which uses negative feedback in its convolution
process to attenuate frequency content above its cut off frequency. Unlike an analog filter, the
advantage of using a digital filter is that the attenuation is precise and will begin at the desired
cut off frequency.

Through this lab, we will be able to design and create various active analog circuits and compare
their real behavior to the ideal behavior. Additionally, we will be able to compare digital and
analog filtering methods.

2
Methods

The frequency response of an inverting amplifier (Fig. 1) was determined for two configurations
of two circuits with input resistance (Ri) and feedback resistance (Rf) of 1 kΩ and 20 kΩ, and 1
kΩ and 100 kΩ, respectively. The 741 op-amp was powered by a DC power supply (±15 V).

Figure 1. Inverting amplifier with non-unity gain

DC offset was added to an input signal by adding a 5 kΩ potentiometer in series with the input
resistor. The range of the DC offset was then determined by moving the wiper on the
potentiometer through its range.

Figure 2. Adding DC Offset

3
To create a low pass, high pass and band pass filter, three different RC circuits were created. The
741 op-amp was used for all of the filters. The low pass filter was created with an Rf of 10 kΩ
and a capacitor of 0.001 µF (Fig. 3). The theoretical cutoff frequency of the low pass filter was
calculated as 1/2πRfC. The actual cutoff frequency was determined by measuring the maximum
gain and determining the frequency at which the gain was 3dB below the maximum. The high
pass filter was created with a Ri of 1 kΩ and a capacitor of 0.1 µF (Fig. 4). The theoretical cutoff
frequency of the high pass filter was calculated as 1/2πRiC. The actual cutoff frequency was
determined by measuring the maximum gain and determining the frequency at which the gain
was 3dB below the maximum. The band pass filter was created with an Rf and Ri of 10 kΩ and 1
kΩ, respectively. And capacitors of 0.1 µF and 0.001 µF, respectively (Fig. 5). The theoretical
bottom cutoff frequency of the band pass filter was calculated as 1/2πRfC2 and the theoretical
top cutoff frequency of the band pass filter was calculated as 1/2πRiC1. The actual cutoff
frequency was determined by measuring the maximum gain and determining the frequency at
which the gain was 3dB below the maximum.

Figure 3. Low Pass Filter

4
Figure 4. High Pass Filter

Figure 5. Band Pass Filter

5
The 3rd and 5th order Butterworth filters were created using LabVIEW software (Fig. 6). To
determine the frequency response of the filters, a series of analog signals with increasing
frequencies were generated by a waveform generator and converted to digital signals using a NI
USB 6009 data acquisition module. The analog signals were sampled at a rate greater than twice
the input frequency as recommended by the Nyquist criterion. After determining the frequency
response of these filters, a noisy signal was generated by adding a 60 Hz sine wave to a 30 Hz
sine wave (Fig. 7 and 8). The 60 Hz signal was removed from the 30 Hz signal but altering the
cutoff frequency of the Butterworth filter and using the frequency spectrum to determine when
the gain of the 60 Hz signal was below 0 dB. The SNR was calculated as the gain of the filtered
signal at 30 Hz divided by the gain of the filtered signal at 60 Hz.

Figure 6. Virtual Instrument for 3rd and 5th order Butterworth Filter

6
Figure 7. Circuit to generate noisy 30 Hz signal

Figure 8. Virtual Instrument to add and remove noise (60 Hz sine wave) from the 30 Hz sine wave

Results

1. Frequency vs. Gain of Unity Gain Follower

7
Frequency Gain
100 Hz 0.97
10 kHz 0.96
100 kHz 0.94
1 MHz 0.92

2. Bandwidth at -3 dB point: 1.39 MHz

3. Phase angle at -3 dB point: 49 ˚

4. Gain and Bandwidth of a Non-Unity Inverting Amplifier

Vi Ri Rf Gain BW (Hz)
400 mVp-p 1 kΩ 20 kΩ 17.3 39 kHz
200 mVp-p 1 kΩ 100 kΩ 8.5 82 kHz

5. Sign of the gain: Negative

6. The gain and the bandwidth are inversely related by a factor of 2. As you increase the
gain, you decrease the bandwidth which allows fewer frequencies to affect the signal
content.

7. The maximum input voltage is 20 Vpp.

8. The range of DC offset is 40.34 V – 46.88 V.

9. To increase the DC offset range available without affecting the gain the values of the
resistors in the circuit can be changed.

10. Gain and Phase of the filter at the theoretical -3dB point
Frequency Gain (dB) Gain (difference Phase (degrees)
from max)
15 kHz 14.27 -3.89 -117

11. Gain and phase table of the filter

Frequency Gain (dB) Gain (difference from Phase (degrees)


max)
50 Hz 17.68 -0.48 -200
100 Hz 17.33 -0.83 -190

8
150 Hz 16.71 -1.45 -190
200 Hz 17.82 -0.34 -190
300 Hz 16.85 -1.31 -188
400 Hz 16.78 -1.38 -185
500 Hz 16.97 -1.19 -183
600 Hz 17.23 -0.93 -174
700 Hz 17.29 -0.87 -160
1 kHz 17.15 -1.01 -165
2 kHz 16.57 -1.59 -157
3 kHz 16.55 -1.61 -152
4 kHz 16.35 -1.81 -147
5 kHz 16.25 -1.91 -146
6 kHz 16.16 -2.00 -140
7 kHz 15.92 -2.24 -139
8 kHz 15.90 -2.26 -139
9 kHz 15.78 -2.38 -136
10 kHz 15.72 -2.44 -136
11 kHz 15.65 -2.51 -136
12 kHz 15.66 -2.50 -136
13 kHz 15.66 -2.50 -136
13.1 kHz 15.32 -2.84 -128
13.2 kHz 15.04 -3.12 -127
14 kHz 14.62 -3.54 -120
15 kHz 14.27 -3.89 -117
15.5 kHz 14.19 -3.97 -115
16 kHz 14.09 -4.07 -111

9
12. Measured -3 dB Point: 13.1 – 13.2 kHz

13. Filter type: Low Pass Filter

14. Step Response of the filter to a 1 Vpp Square wave with a frequency equal to the
theoretical cut off frequency

10
15. The output signal decreases in response to the input signal.

16. Gain and Phase of the filter at the theoretical -3dB point
Frequency Gain (dB) Gain (difference Phase (degrees)
from max)
1.59 kHz 17.34 2.66 -117

17. Gain and phase table of the filter

Frequency Gain (dB) Gain (difference Phase (degrees)


from max)
100 Hz 0 -20 -60
150 Hz 0.2593 -19.7407 -62
200 Hz 2.175462 -17.8245 -63
300 Hz 6.597735 -13.4023 -85
400 Hz 7.130946 -12.8691 -85
500 Hz 8.952556 -11.0474 -86
600 Hz 10.05351 -9.94649 -85
700 Hz 11.39244 -8.60756 -90
800 kHz 12.07404 -7.92596 -93
900 kHz 13.43636 -6.56364 -95
1 kHz 14.11002 -5.88998 -96

11
1.25 kHz 15.2998 -4.7002 -97
1.5 kHz 16.54 -3.46 -99
1.6 kHz 17.01562 -2.98438 -103
2 kHz 19.00814 -0.99186 -105
3 kHz 22.42087 2.420874 -107
4 kHz 25.54732 5.547322 -111
5 kHz 27.69152 7.691521 -111

18. Measured -3 dB Point: 1.6 kHz

19. Filter type: High Pass Filter

20. Step Response of the filter to a 1 Vpp Square wave with a frequency equal to the
theoretical cut off frequency

12
21. The output signal was unstable and oscillates in response to the input signal

22. Gain and Phase of the filter at the theoretical -3dB point
Frequency Gain (dB) Gain (difference Phase (degrees)
from max)
15 kHz 29.28 -1.79 -190
1.59 kHz 17.59 -13.48 -110

23. Gain and phase table of the filter

Frequency Gain (dB) Gain (difference from Phase


max) (degrees)
150 0.00 -31.07 -74
200 1.49 -29.57 -74
300 4.57 -26.50 -74
400 6.80 -24.27 -74
500 8.49 -22.58 -83
600 9.72 -21.35 -84
700 10.98 -20.09 -85
800 12.24 -18.83 -87
900 12.86 -18.21 -89
1000 13.98 -17.09 -94
1250 15.65 -15.42 -101

13
1500 17.23 -13.83 -108
1590 17.59 -13.48 -110
1600 18.76 -12.31 -112
2000 19.87 -11.20 -115
3000 22.74 -8.32 -120
4000 25.61 -5.46 -130
5000 27.96 -3.11 -140
10000 31.07 0.00 -190
20000 29.45 -1.62 -220
20500 28.45 -2.62 -221
21000 27.94 -3.12 -223
21500 27.38 -3.69 -225
22500 27.15 -3.92 -228
25000 26.68 -4.38 -230
30000 24.82 -6.25 -234
40000 22.61 -8.46 -245
50000 21.34 -9.73 -250
60000 20.00 -11.07 -251
70000 18.42 -12.65 -276
80000 18.21 -12.86 -283
90000 17.45 -13.62 -283
100000 16.62 -14.44 -283

14
24. Measured -3 dB Point: 5 kHz and 21 kHz

25. Filter type: Band Pass Filter

26. Step Response of the filter to a 1 Vpp Square wave with a frequency equal to the
theoretical cut off frequency

15
27. The output signal decreases in response to the input signal.

31. Frequency response of a 3rd and 5th order Butterworth filter

3rd order Butterworth Filter


Frequency (Hz) Gain Phase
5 1 -30.6
40 0.94 -115.2
50 0.72 -126
60 0.52 -172.8
500 0.21 -180

16
5th order Butterworth Filter
Frequency (Hz) Gain Phase
5 1 -97.2
40 0.97 -172.8
50 0.74 -216
60 0.38 -280
500 0.19 -360

32. Filtering a noisy signal

a. Cutoff frequency: 35 Hz. This was the frequency at which the signal became a sine
wave with a frequency of 30 Hz.

b. Signal to Noise Ratio (SNR) at cutoff frequency: 31.6. The SNR was computed as the
gain (Vout/Vin) of the frequency spectrum at 30 Hz divided by the frequency at 60 Hz.

c. Graph of filtered and unfiltered signals

Discussion

Through a series of experiments in this lab, we were able to design a unity and non-unity gain
amplifier, add DC offset to a signal and create three RC circuits to perform high pass, low pass
and band pass filtering. Additionally, we were able to determine the frequency response of a
couple of digital low pass filters and determine the cutoff frequency to remove noisy frequency
content from a signal. The measured gain of the unity gain amplifier was less than one and
decreased as the frequency was increased. The gain and the bandwidth of the inverting amplifier
were determined to be inversely related by a factor of 2. As the gain was increased, the
17
bandwidth was decreased which allows fewer frequencies to affect the signal content. A
doubling of the gain reduced the bandwidth of the signal by half. The relationship between the
gain and bandwidth of the inverting amplifier highlights an important point in the selection of
amplifier components. The range of DC offset was 40.34 V – 46.88 V. The measured cutoff
frequencies of the active filters were not equal to the ideal cutoff frequencies. The measured
cutoff frequencies were 13.1 kHz, 1.6 kHz and 5 kHz and 20 kHz for the low pass, high pass and
band pass filters, respectively. The cutoff frequency to eliminate 60 Hz noise from a 30 Hz signal
was 35 Hz and the signal to noise ratio was 31.6.

For the analog filters, the frequency response had a non-uniform gain after the cutoff frequency.
After the cutoff frequency, the low pass filter continued to attenuate high frequency content
while the sign of the gain became more positive (in phase). As for the high pass filter, the
frequency content was amplified after the cutoff frequency while the sign of the gain became
more negative (out of phase). Finally, for the band pass filter, the frequency content was
amplified before the bottom cutoff frequency and attenuated after the upper cutoff frequency. In
between both cutoff frequencies, the gain was a fraction of a unity gain. Although we did not
determine the frequency response for many more frequencies after the cutoff frequency, the gain
of the filters began to saturate at frequencies exceeding the cutoff frequency. Unlike a passive
filter which has an infinite bandwidth, the limited bandwidth of an active filter results in gain
saturation where changes in frequency are not accompanied by gain increases.

The digital low pass filters were more precise and accurate in comparison to the analog low pass
filters. However, to adequately recreate a 30 Hz sine wave from a sine wave which was created
by summing a 60 Hz and 30 Hz sine wave, the ideal cutoff frequency was 35 Hz with a SNR of
31.6. This indicates that regardless of analog or digital filtering, the cutoff frequency needs to be
closer to the frequency of the signal as opposed to noise.

The cutoff frequency step response of the high pass filter was very unusual. As opposed to
exponentially attenuating the input signal in the manner of a first order response, the output
signal oscillated before (and while) attenuating. A possible explanation for this response is that
the measured cutoff frequency was lower than it actually was. The error most likely resulting
from the measurement process and the incorrect determination of the maximum gain.

Through this lab, we learned how to design and create various active analog circuits and compare
their real behavior to the ideal behavior. Additionally, we will be able to compare digital and
analog filtering methods. Overall, this process gave us an appreciation for digital filtering and the
tools to troubleshoot active analog circuits.

18
Simisola O. Oludare

Bioe 431: Instrumentation Lab I

Lab partner: Nadia Crawley

Section 14163

TA: Yanjun Chen

Thermistor Linearization and Modeling

Abstract

A patient’s body temperature is an indication of her health status. For this reason, a clinician
needs to be able to accurately measure, display and record changes in body temperature. To
achieve this, an engineer can develop a device which uses a temperature transducer, an electronic
display unit and an analog-to-digital converter to create equipment capable to measuring,
displaying and recording temperature changes as voltage changes. In this lab, we created an
electronic circuit with a negative temperature coefficient thermistor capable of performing those
functions. Additionally, we were able to characterize the frequency response of the thermistor.
To create the electronic circuit, we calibrated the thermistor and determine the temperature
coefficient to be 0.002%/K and beta (material property to convert temperature to resistance) as
4535.41 K. The gain of the thermistor circuit was attenuated from 141mV /˚C to 1mV/˚C using a
resistor (Rf) of 2.12 kΩ and resistor (Ri) of 300 kΩ. Then the voltage change was made equal to
and one-to-one to the temperature change using a 20 kΩ potentiometer. Finally, the time constant
and bandwidth of the thermistor were determined to be 8.67 seconds and 0.02 Hz, respectively.
The thermistor was identified as a first order system which is analogous to an RC circuit.
Overall, this lab allowed us to learn how to work with nonlinear transducers through the use of
thermistors.

Keywords

Temperature transduction, negative temperature coefficient thermistor, calibration, linearization,


characterization

Introduction

In medicine, a patient’s body temperature is an indication of health. An extreme drop in external


body heat is an early indicator for shock due to low blood flow to the periphery (Peura and
Webster, 1998). Conversely, an increase in temperature is indicative of an infection due to
increased perspiration and blood flow (Peura and Webster, 1998). For these reasons it is
necessary for clinicians to be able to accurately measure, display and record external
temperature. To develop equipment capable of accurately measuring, displaying and recording

1
temperature, engineers need to create electronic devices capable of responding to changes in
temperature. Such a device would need to contain a temperature transducer which transforms
changes in temperature to changes in voltage. Additionally, the changes in voltage will need to
be calibrated such that there is a one-to-one relationship between change in temperature and
voltage.

Three common types of temperature transducers are thermocouples, resistance temperature


detectors (RTDs), and thermistors. A thermocouple works by exploiting the Seebeck effect
which states that a continuous current flows between two dissimilar metals when one junction is
heated. By breaking the loop of the metals, the voltage across the two terminals of the metals can
be measured. An RTD is a thin wire or film which has a change in resistance if it is heated. And
a thermistor is a ceramic semiconductor which has dramatic changes in resistance due to a
change in temperature.

In this lab, a negative temperature coefficient (NTC) thermistor was used as the temperature
transducer. As indicated by its name, the NTC thermistor has an inverse nonlinear relationship
between temperature and resistance. Although the thermistor is better than a thermocouple and
RTD at detecting minute changes in temperature, it is difficult to work with due its nonlinear
nature. However, these difficulties can be avoided by calibrating and linearizing the thermistor.
In addition to the thermistor circuit, a potentiometer will be used to adjust the associated change
in voltage such that there is an equal and one-to-one change in voltage and temperature. Finally,
a step response will be used to characterize and model the thermistor by determining its time
constant and bandwidth.

Methods

The thermistor was calibrated by placing it in series with a known resistance (30 kΩ) and
measuring the potential difference across the thermistor in response to a 5 V sine wave (Fig. 1).
The potential difference and known resistance were then used to calculate the resistance of the
thermistor by the formula: . The resistance of the thermistor was then determined
every 10 ˚C over 0 ˚C to 100 ˚C and every 5˚C over 20˚C to 40˚C.

Figure 1

2
The thermistor was linearized with an equivalent resistance (Rn) which is created by placing the
thermistor in parallel with a resistor equal to the resistance of the thermistor at 37˚C (Fig. 2). The
equivalent resistor (Rn) was then placed in series with a resistor of equal resistance (Fig. 2).
These resistors were plugged into the inverting input of a 741 op-amp which was powered by a
15 V direct current (DC) with a voltage follower. The equivalent resistance and output voltage
was then calculated over the range: 0˚C to 50˚C. By placing the resistors in this configuration,
the equivalent resistance will be equal to the thermistor’s 37˚C resistance at low temperatures
and equal to RT at high temperatures.

Figure 2

A circuit with a voltage divider and a 20 kΩ potentiometer was then built to adjust the gain and
offset of the thermistor circuit to 1 mV/˚C and 0 ˚C, respectively (Fig. 3). The Rf used in the
voltage divider was determined by determining the gain (-Rf/Ri) necessary to attenuate the slope
of the V/T relationship developed for the linearized circuit (Fig. 2). The output voltage of the
circuit was displayed using a commercially available digital panel meter.

Figure 3

The frequency response was determined by observing the thermistors response to a step input. A
step input was simulated by measuring ambient temperature then plunging the thermistor into a

3
tub of 60˚C. To measure the step response, the thermistor was placed in series to a 30kΩ resistor
across which the potential difference was measured (Fig 4).

Figure 4

Results

Q1)

Rn = 2.51 kΩ

Q2)

T (˚C) Vi Rn
10 3.44 2.45
20 2.4 2.4
30 1.18 2.33
37 0.252 2.28
40 -0.4 2.25
50 -0.2 2.12

2.5 y = -0.0081x + 2.5613


2.45 R² = 0.9532
2.4
2.35
Rn

2.3
2.25
2.2
2.15
2.1
0 10 20 30 40 50 60
Temperature (˚C)

4
Q3)

5
4 y = -0.1407x + 5.0358
3 R² = 0.9932
2
1
Voltage

0
-1
-2
-3
-4
-5
0 10 20 30 40 50 60 70
Temperature (˚C)

Q4)

Q5)

70
Unadjusted
60

50
y = 1.0023x + 13.319
40
Voltage

R² = 0.9969

30

20

10

0
29 34 39 44 49
Temperature (˚C)

5
Q6)

50
45
Adjusted
40
35 y = 1.0023x - 0.1814
30 R² = 0.9969
Voltage

25
20
15
10
5
0
29 34 39 44 49
Temperature (˚C)

Q7)

160000
140000
120000
RT (Ohms)

100000
80000
60000
40000
20000
0
273.15 293.15 313.15 333.15 353.15 373.15
Temperature (Kelvins)

6
Q8)

Temp (˚C) Temp (K) Beta


35 308.15 4535.234
37 310.15 4451.83
40 313.15 4619.166
Average 4535.41

Q9)

200000
180000
160000
140000
120000 R-squared = 0.98
RT (Ohms)

Theoretical
100000
Observed
80000
60000
40000
20000
0
273.15 293.15 313.15 333.15 353.15 373.15
Temperature (Kelvins)

Q10)

The temperature coefficient is found by taking the slope of the curve in the range of 308.15 K -
313.15 K: -β/T2. α=-0.002%/K

Q12) The thermistor appears to be a first order system.

7
Step Response
7

4
Voltage

0
0 10 20 30 40 50 60 70
Time (seconds)

Q13)

An RC circuit in series is mathematically equivalent to the thermistor. The time constant is


equivalent to the multiplication of R and C.

Q14)

The time constant of the thermistor, τ, is equal to 8.67 seconds. The bandwidth of the thermistor,
f3db= 0.02 Hz.

Discussion

Through this lab, we have learned specifically how to calibrate, linearize and characterize a
nonlinear temperature transducer. By calibrating the thermistor, we were able determine the
relationship between resistance change and temperature change. As expected, the relationship
was inverse and nonlinear. The material constant, beta, which relates the change in temperature
exponentially to a baseline temperature, was computed to be 4531.41K, a 13.3% error from the
typically expected value of 4000K. The error is probably due to changes in the thermistor
properties after exposure to extreme temperatures and constant use. Additionally, the
temperature coefficient was determined to be 0.002% decrease in resistance as temperature (in
Kelvins) increases.

8
Based on the calibration data, we were able to determine the resistance of the thermistor at 37˚C
and use it to linearize the thermistor. The thermistor was linearized by creating an equivalent
resistance of . The linearity achieved using a Req of 2.5 kΩ was r = 0.99 and the
measured Rn at 37˚C was calculated to be 2.28 kΩ. Although the measured Rn at 37˚C has an
error of 9.2%, the discrepancy is small and can be explained by the difference in resistance
values used. Additionally, by using a voltage divider with a gain of 0.007, we were able to
generate a 1mV/˚C from our circuit.

Finally, the frequency response of the thermistor was determined using a step input which
involved a rapid change in temperature from ambient temperature to 60˚C. As expected, there
was an increase in voltage when the thermistor was plunged into the hot water and the settling
time was approximately 30 seconds. The time constant and the bandwidth of the thermistor were
8.67 seconds and 0.02 Hz, respectively. This means that the thermistor functions like a low pass
filter, an RC series circuit, which amplifies low frequencies.

Overall, this lab has given us the experience and skills necessary to work with nonlinear, variable
sensors such as strain gauges and potentiometers. These are sensors which are commonly used to
measure bio signals which inform health status. With the knowledge and skills from this lab, we
will be able to make accurate and meaningful measurements from nonlinear, variable sensors.

9
References

Peura R.A. and Webster J.G. Chapter 2: Basic Sensors and Principles. Webster, J. G. (Ed).Medical
Instrumentation: Application and Design (233-286). Massachusetts, USA: John Wiley &
Sons

10
Simisola O. Oludare

Bioe 431: Instrumentation Lab I

Lab partner: Nadia Crawley

Section 14163

TA: Yanjun Chen

Biopotential Amplifiers and the Electrocardiogram

Abstract

The purpose of this lab was to design and build a biopotential amplifier to record the electrical
activity of the heart (electrocardiogram). The biopotential amplifier consisted of a differential
amplifier, a band pass filter and a voltage divider. Two biopotential amplifiers were built for this
lab. The first consisted of a three operational amplifier differential amplifier and the other
consisted of the commercially available AD620 differential amplifier. The electrocardiogram
was collected by placing electrodes on the skin in Eindhoven’s triangle and connecting the
electrodes to the biopotential amplifier with a ground electrode placed on the right ankle. Before
recoding the electrocardiogram, the specifications of the amplifier were measured and calculated
using artificially generated common mode and differential mode signals. The common mode
rejection ration of the biopotential amplifier was 3.67, the slew rate was 30.4 V/sec and the
measured -3dB point was 1000 Hz. With the three operational differential amplifier biopotential
amplifier, we were able to record one period of the electrocardiogram wave form. However, the
signal had 60 Hz interference due to the wide range of the band pass filter. The
electrocardiogram measured from the biopotential amplifier with the AD260 differential
amplifier also had 60 Hz interference and a DC shift from the power line. Through this lab, we
were able to learn how to precisely and accurately measure and record physiological electrical
activity from the heart with the understanding that the principles learned here can be applied to
other physiologically generated electrical signals.

Keywords

Heart, Electrical Activity, Surface Electrical Activity, Eindhoven’s triangle, Electrocardiogram,


Biopotential Amplifier, Differential Amplifier, Band Pass Filter

Introduction

The functions of many organs and organ systems in the human body are mediated by excitable
cells which generate electrical signals. One of those organs is the heart. The heart requires the
generation of electrical signals to stimulate muscles required for the circulation of blood
throughout the cardiovascular system. The heart’s electrophysiology has been widely studied and
the structure of the signal is often used to predict heart disease states such as arrhythmias.
The electrical signal of the heart is generated when the pacemaking cells in the sinoatrial (SA)
node (located at the junction of the superior vena cava and the right atrium) are depolarized. This
electrical impulse is then transmitted throughout the heart when other cells and tissue in the heart
are depolarized. The pathway of the electrical impulse is from the SA node to the right and then
left atria via special conducting tracts. Then from the atria, the impulse is transmitted to the right
and left ventricle via the atrioventricular node into the bundle of His, the right bundle branch, the
anterior and posterior divisions of the left bundle branch and then into the left branch. The
overall wave form generated by the activation of these nodes and bundles are called the PQRS
and T waves (Webster et al, 1998). The P wave is generated by atrial depolarization, the QRS
wave is generated by ventricular depolarization and the T wave is generated by atrial
repolarization. The magnitude of the QRS wave is typically 1-3 mV and lasts approximately
200-300 ms (Webster et al, 1998). Overall, the frequency of the heart’s electrical activity is in
the range of 0.3 Hz – 120 Hz.

To capture this wave form, the heart is modeled as an electrical equivalent generator where the
electrical activity at each region of the body is represented by a current dipole and a net dipolar
contribution from all active areas relative to the heart (Webster et al, 1998). The electrical
activity at each region of interest is captured by placing electrodes on the surface of the skin in a
configuration known as Eindhoven’s triangle in which the electrodes are placed at the right wrist,
left wrists and right ankle. By placing electrodes in this configuration, it is possible to measure
the potential difference of the heart based on Kirchhoff’s voltage law.

The circuit which is used to collect the heart’s electrical activity is called an electrocardiograph.
An electrocardiograph is a biopotential amplifier made up of three functional blocks which
collect the heart electrical activity and condition it so that the signal is accurate and useful. The
first block is a differential amplifier which is used to amplify potential differences across the
right and left wrist electrodes. The next block is a band pass filter which is used to filter
interfering electrical activity from the electrical sources intrinsic and extrinsic to the body such
as skeletal muscle and the power line (60 Hz).

The purpose of this lab was to design an electrocardiograph (ECG) and use it to accurately
record the electrical activity of the heart. In this lab, we will be building two biopotential
amplifiers and comparing their ability to accurately collect the heart’s electrical activity.

Methods

Two biopotential amplifiers were constructed in this lab. The first was a made using a three
operational amplifier (op amp) differential amplifier and the
Figure 1. Schematic for a three OP Amp Biopotential Amplifier

The specs

The frequency response

The Biopotential amp with the commercially available differential amp

Results

The Biopotential amp with the three op amp differential amp


The specs

Common mode gain (at 50 Hz) 0.89


Differential mode gain (at 50 Hz) 3.25
Common mode rejection ratio 3.67
Slew rate at 10 Hz (V/secs) 30.4
The frequency response

Frequency Gain dB
(Hz)
0.3 0
1 -0.5891
10 -0.56873
50 -0.24129
75 -0.47606
125 -1.09478
150 -1.358
200 -1.20832
250 -0.96609
300 -0.97973
1000 -2.80902
2000 -4.55202
10000 -9.00938
The Biopotential amp with the commercially available differential amp

Discussion

The purpose of this lab was to create two Biopotential amps. The specs. The signals for both.
The frequency response.

Things to be aware of and how they were accounted for: electrode placement, interference
sources, shocking the patient.

The principles used in this lab can be applied to other sources of biopotnetials

Through this lab we learned how to build a Biopotential amp and use it. We also learned that the
ideal signal is tough to acquire because of the environment. Intrisic to the person and extrinsic to
the person.
Simisola O. Oludare

Bioe 431: Instrumentation Lab I

Lab partners: Karolina Blaszczuk and Victoria Perizes

Section 14163

TA: Yanjun Chen

Force measurements using strain gages and Wheatstone bridge

Abstract

A strain gage is an indirect force transducer which has a change in resistance that is dependent
on deformation. . By using the equations of the beam theory, the change in deformation can then
be used to determine the applied force. Thus, a simple and practical way to determine the relation
of deformation to resistance change is to use known loads to deform the strain gage and measure
the voltage across a circuit. In this lab, we characterized the voltage-force relationship of a single
strain gage and a strain gage in a Wheatstone bridge circuit (with and without a differential
amplifier) by using known loads to determine the voltage difference. The Wheatstone bridge and
differential amplifier proved to be the most sensitive strain gage configuration with sensitivity of
0.000003 V/mN. The Wheatstone bridge without the differential amplifier was the least sensitive
with a value of 0.000000008 V/mN. The single strain gage had a sensitivity of 0.000001 V/mN.
Overall, the Wheatstone bridge was able to provide two functions to the strain gage: linearization
and increased sensitivity. Through this lab, we learned how to implement a force transducer into
a differential amplifier and also how to use a Wheatstone bridge to increase the functionality of
the transducer

Keywords

Beam theory, strain, variable resistance, strain gage, Wheatstone bridge, differential amplifier

Introduction

The human body generates forces and motions which can be used to determine its underlying
function. For example, grip force is used as a measure of healthy aging in older adults. To
measure and record grip force, bioengineers need to be able to transduce the applied force into an
electrical signal. One method of doing this is with a strain gage. A strain gage is an indirect force
transducer which has a change in resistance that is dependent on deformation. Strain gages are
typically made of thin metal films which are bonded together in a zigzag pattern. In a zigzag
pattern, the stretching of the strain gage leads to lower resistance. By using the equations of the
beam theory, the change in deformation can then be used to determine the applied force. The rate
at which resistance changes due to deformation change is called the gage factor. To determine
the gage factor, it is necessary to deform the strain gage at known distances and measure the
change in voltage across a circuit. However, it is difficult to measure deformation accurately in a
laboratory setting. A simpler and more practical way is to use known loads to deform the strain
gage and measuring the voltage across a circuit. From the equations of the beam theory, it is then
possible to calculate the deformation of the strain gage and the resulting resistance changes.

In this lab, we characterized the force-voltage relationship of a single strain gage and a strain
gage in a Wheatstone bridge circuit by using known loads to determine the voltage difference. A
Wheatstone bridge circuit is a quadrilateral configuration of four resistors in which the voltage
change across the circuit is dependent on the difference between the ratios of adjacent resistors.
If these ratios are equal, there is no overall resistance difference and thereby no voltage
difference across the circuit. However, if these values are different, there is an overall resistance
difference resulting in a voltage difference across the circuit. For strain gages in a Wheatstone
bridge circuit, the strain gages on one adjacent pair have resistance changes due to tension and
compression for the other pair. If there is no difference between these ratios, the circuit is
balanced and there is no voltage difference across the circuit. However, if there is a difference
between these ratios, the circuit is unbalanced and will result in a voltage difference. This
voltage difference can then be amplified using a differential amplifier. Theoretically, this means
that the Wheatstone bridge circuit will be more sensitive to changes in deformation than the
single strain gage. Through this lab, we will to test this prediction by comparing the sensitivity of
a Wheatstone bridge circuit with that of a single strain gage.

Methods

The voltage-force relationship of a single strain gage was determined by creating a circuit which
consisted of a voltage divider, voltage follower and a differential amplifier. The strain gage used
in this lab was made of foil (CEA-06-187UW-120) bonded to a phosphor bronze bar. Typically,
the accuracy of the strain gage is affected by changes in temperature because increased
temperature leads to volume expansion which affects resistance. However, the normal range for
this strain gage is -100 ˚F to 350 ˚F the normal temperature range so we did not have to account
for temperature. The loads which were used to determine the voltage-force relationship were 0g,
5g, 15g, 25g, and 35g.

The voltage divider was created by placing the strain gage in series with a known resistor (120
Ω) and powered by 5 V DC. The voltage difference across the strain gage was then connected to
a voltage follower created using a 741 operational amplifier (op amp). The output from the
voltage follower was then connected to the non-inverting input of the differential amplifier. And
a 20 kΩ potentiometer with a series of resistors (620 kΩ and 1.2 MΩ) was connected to the
inverting input of the differential amplifier.
Figure 1. Single strain gage circuit

The voltage-force relationship of the Wheatstone bridge circuit was determined by measuring the
voltage difference between adjacent strain gages. The voltage-force relationship was measured
with and without connection to the differential amplifier. The loads which were used to
determine the voltage-force relationship of the Wheatstone bridge without the differential
amplifier were 0g, 5g, 15g, and 25g. And the loads which were used to determine the voltage-
force relationship of the Wheatstone bridge with the differential amplifier were 0g, 10g, 20g,
30g, 35g, 50g, and 60g.

The output of the Wheatstone bridge circuit without the differential amplifier was measured after
the voltage followers in pseudo-differential mode (Figure 2). For the Wheatstone bridge circuit
with the differential amplifier, the input signal connected to the non-inverting input was placed
in series with a 20 kΩ potentiometer and a series of resistors (Figure 2). The potentiometer was
used to shift the max voltage of the differential amplifier.

Figure 2. Wheatstone bridge circuit

Results

Single strain gage

1. Differential gain = R4/R3 = 100


0.6
y = 1E-06x + 0.1502

0.5

0.4
Voltage (V)
0.3

0.2

0.1

0
0 50000 100000 150000 200000 250000 300000 350000 400000
Force (milliNewton)

Figure 3. Voltage measurements for a single strain gage.

3. The sensitivity of the strain gage is 0.000001 V/mN.

Wheatstone bridge without differential amplification

0.16
y = 8E-08x + 0.1166
0.14
0.12
0.1
Voltage (V)

0.08
0.06
0.04
0.02
0
0 50000 100000 150000 200000 250000 300000
Force (milliNewton)

Figure 4. Voltage measurements of Wheatstone bridge without differential amplification

2. The sensitivity of the Wheatstone bridge without amplification is 0.000000008 V/mN


Wheatstone bridge with differential amplification

1.4 y = 3E-06x + 0.2623 1.6 y = 2E-06x + 0.4275


1.2 1.4
1 1.2

Voltage (V)
Voltage (V)

1
0.8
0.8
0.6
0.6
0.4 0.4
0.2 0.2
0 0
-50000 50000 150000 250000 350000 0 200000 400000 600000
Force (milliNewton) Force (milliNewton)

Figure 5. Voltage measurements of Wheatstone bridge with differential amplification

2. The sensitivity of the Wheatstone bridge with amplification before it saturates is


0.000003 V/mN and 0.000002 V/mN after it saturates.
3. At 0 N, the bridge drifts between 60 mV and 120 mV. The measurement is linear until
343,000 mN after which it asymptotes.

Discussion

The purpose of this lab was to characterize a single strain gage and a Wheatstone bridge circuit
of strain gages. The strain gage serves as force transducer by transforming deformation to
resistance changes. Consequently, this relationship can be characterized by determining the
voltage-force relationship because through the beam theory, it is possible to determine the
deformation-resistance relationship. To determine the voltage-force relationship, known loads
were used to determine the voltage difference across the circuit. Due to the small resistance
changes of the strain gage, a differential amplifier was necessary to amplify the voltage
difference created in both the single strain gage circuit and the Wheatstone bridge circuit.

As predicted, the Wheatstone bridge and differential amplifier proved to be the most sensitive
strain gage configuration with sensitivity of 0.000003 V/mN. However, the Wheatstone bridge
without the differential amplifier was the least sensitive with a value of 0.000000008 V/mN. This
low sensitivity value is due to the low resistance changes which the strain gage produces. The
single strain gage had a sensitivity of 0.000001 V/mN which was slightly lower than that of the
Wheatstone bridge circuit. Additionally, with the Wheatstone bridge circuit, the voltage
difference across the circuit was linearized unlike the single strain gage circuit which was
logarithmic.
Overall, the Wheatstone bridge was able to provide two functions to the strain gage: linearization
and increased sensitivity. Through this lab, we learned how to implement a force transducer into
a differential amplifier and also how to use a Wheatstone bridge to increase the functionality of
the transducer. This technique can also be applied to any other circuit with a variable resistor
sensor.
Simisola O. Oludare

Bioe 431: Instrumentation Lab I

Lab partners: Karolina Blaszczuk and Victoria Perizes

Section 14163

TA: Yanjun Chen

PHOTOPLETHYSMOGRAPHER PCB CIRCUIT

Abstract

The purpose of this lab was to build a photoplethysmographer and use it to measure
blood flow. The photoplethysmographer was built using a light emitting diode (LED) as the light
source and a phototransistor as the light collector. A printed circuit board (PCB) was used to
build the photoplethysmographer. After the photoplethysmographer was built, it was tested by
having a subject place their fingertip on the HOA 1397-32. The circuit was built to specifications
and we observed the LED flickering due to the changes in blood flow associated with the cardiac
cycle. Through this lab, we learned soldering techniques which will are useful for working with
PCB.

Keywords

Photoplethysmography, printed circuit board, soldering, light emitting diode, phototransistor

Introduction

Plethysmography is any method used to measure the blood flow through the body. One
method of plethysmography is the use of optoelectronics to create a photoplethysmographer. A
photoplethysmographer works by using an infrared (IR) emitter to penetrate the skin and muscles
at the fingertip. This light is then absorbed by the blood and reflected by the bone. The reflected
light is then collected using an IR detector. The amount of light collected by the detector is a
function of the distance of the bone and the density of blood at the fingertip. Based on these
parameters, the amount of blood flowing through the body can be calculated from the amount of
light collected. By building a circuit around the light source, the variation in the light collected
can be converted to voltage allowing us to quantify, analyze and record blood flow. In this lab,
we will be building such a circuit using a printed circuit board (PCB). A PCB will be used to
build the photoplethysmographer because it is more robust to bumping and short circuits which
are typically problems that occur when using a breadboard. However, the PCB is also prone to
leakage resistances and environmental factors which lead to parasitic effects. A PCB is
composed of a copper layer containing the signal traces, joints, powers and grounds which are
sandwiched in between two insulating layers. Circuit components are attached to the PCB
through soldering which is a method of joining two metals by melting solder (the filler material)
onto both materials; in this case, the joint and the circuit components. This technique is
performed using a soldering iron, solder and solder wick.

The circuit which we will use to build the photoplethysmographer will consist of a light
emitting diode (LED) as the light source and a phototransistor as the light detector. An LED is a
two-terminal electronic device that has a high resistance to current flow in one direction, and a
low resistance to current flow in the opposite direction. An LED emits light in proportion to the
forward current; it is powered by a voltage source, typically a 9V battery, and needs to be placed
in series with a resistor in order to limit the forward current. A phototransistor is a made up of an
emitter, collector and base. The current flows into the base then the collector and the emitter. In
this lab, the phototransistor will act as an amplifier which means that the current from the base to
the collector will be amplified by a gain (G) of 100. In this lab, the phototransistor which we
used was a Honeywell HOA 1397-32 which consists of two transistors arranged in a Darlington
configuration. A Darlington pair acts as a single transistor which amplifies the current with a
squared gain (G2). Additionally, the HOA 1397-32 is highly susceptible to instability caused by
feedback of the amplifiers output. To limit feedback, a zener diode will be used to limit the
voltage change across the HOA 1397-32. The voltage produced by the zener diode will also be
used to ground the output amplifier.

After the photoplethysmographer is built, it will be tested by having a subject place their
fingertip on the HOA 1397-32. If the circuit was built to specifications, we expect to observe the
LED flickering due to the changes in blood flow associated with the cardiac cycle.

Methods
To create the circuit above, we used a PCB, a soldering iron, solder wick and solder. The
soldering iron was heated to 700 degrees. After each component was placed through the joints of
the PCB, the soldering iron was then used to heat the solder and completely seal the joint.

The photoplethysmographer was tested by having a subject place a finger on the


phototransistor while being as still as possible.

Results

As predicted, the LED flickered when a subject rested her finger on the phototransistor reflecting
the cardiac cycle.

Discussion

The purpose of this lab was to develop a photoplethysmographer which can be used to
measure blood flow through the body. The photoplethysmographer works by using an infrared
(IR) emitter to penetrate the skin and muscles at the fingertip. This light is then absorbed by the
blood and reflected by the bone. The reflected light is then collected using an IR detector.

The major components of the photoplethysmographer were the light source, the light
collector, a zener diode with a voltage follower, a non-inverting amplifier, a high pass filter, and
a 9Vpower source. The power source was used to power the LED, which served as the light
source, and the 741 integrated circuits used in the voltage follower and the inverting amplifier.
The light source was placed in series with a resistor in order to reduce the forward current going
through it. The light source was connected to the HOA 1397-32 phototransistor, which consists
of a base, collector and emitter. The voltage change which resulted from altering the amount of
light sensed by the phototransistor was amplified with the non-inverting amplifier and filtered
with a high pass filter. A high pass filter was used to eliminate baseline values with a cutoff
frequency of 0.3Hz. This voltage difference was fed to the LED, changing the state and
brightness of the light. To limit feedback, a zener diode was used to limit the voltage change
across the HOA 1397-32. The voltage produced by the zener diode will also be used to ground
the output amplifier.
Through this lab, we learned soldering techniques which will are useful for working with
PCB in future projects. Additionally, we learned concepts of Photoplethysmography which can
be used to measure fluid and air flow in other systems.
Simisola O. Oludare

Bioe 431: Instrumentation Lab I

Lab partner: Karolina Blaszczuk

Section 14163

TA: Yanjun Chen

FLOWRATE MEASURES USING THERMODILUTION

Abstract

The purpose of this lab was to determine flowrate using the thermodilution method. The method
required the injection of a small volume of ice cold water to cool down the flowing water. The
change in temperature from baseline levels was measured using a thermistor and the Stewart-
Hamilton equation was used to calculate the flowrate. To determine the efficacy of the
thermodilution method, it was compared to the direct flow rate. The direct flow rate was
measured by determining the amount of time the flowing water filled a 100 mL container. The
average (standard deviation) direct flowrate of the flowing water was 3.52 (0.89) mL/secs. The
flow rate calculated using the thermodilution method was 2.59 mL/secs, 1.75 mL/secs, and 0.88
mL/secs when a fast, half of the fast and a quarter of the fast ice water injection speeds were
used. The accuracy of the flow measurement was directly related to injection speed. A possible
source of error is the change in baseline temperature of the flowing water between the fast and
slow conditions. However, the Stewart-Hamilton equation is dependent on the change in
temperature so this is not likely. Additionally, the flowrate of the flowing water was calculated
when the flowing water was pinched and the fastest injection speed was used. The average
(standard deviation) of the pinched rate was 1.81(0.47) mL/secs. Through this lab, we were able
to combine instrumentation and analytical tools to develop a method of determining flowrate.

Keywords

Indicator dilution, flowrate, blood flow, Stewart-Hamilton equation

Introduction

In medicine, the rate of blood flow is an important measure of a patient’s health. Through the
blood flow rate, it is possible to ascertain the supply of oxygen and nutrients to the body’s
organs. For this reason, blood flow needs to be measured.

There are many invasive and noninvasive ways to measure blood flow . The noninvasive
methods include Doppler and electromagnetic flow meters. The invasive methods are those
which use mechanical transducers and indicator dilution technique s. In this lab, we will be using
an indication dilution method to determine blood flow, specifically thermodilution.
Generally, the indication dilution method involves injecting a substance into the blood vessel
and estimating flow rate by monitoring that substance. For thermodilution, the injected substance
is ice cold water and temperature change is measured. Then using the Stewart-Hamilton equation
(below), the flow rate can then be calculated. Where the subscript “b” is in relation to blood and
the subscript “i” is in relation to the injected substance. In the equation, T is the temperature, V is
the volume, c is the specific heat, ρ is the density.

In this lab, the cardiovascular system will be simulated using an aquarium pump and Tygon
tubing ejecting water at room temperature. The injected substance will also be ice cold water.
Through this lab, we will be able to determine the accuracy of the thermodilution method relative
to direct measurements.

Methods

Thermistor

The thermistor was calibrated by placing it in series with a known resistance (30 kΩ) and
measuring the potential difference across the thermistor in response to a 5 V sine wave (Fig. 1).
The potential difference and known resistance were then used to calculate the resistance of the
thermistor by the formula: . The resistance of the thermistor was then determined
every 10 ˚C over 0 ˚C to 100 ˚C and every 5˚C over 20˚C to 40˚C.

Figure 1

The thermistor was linearized with an equivalent resistance (Rn ) which is created by placing the
thermistor in parallel with a resistor equal to the resistance of the thermistor at 37˚C (Fig. 2). The
equivalent resistor (Rn ) was then placed in series with a resistor of equal resistance (Fig. 2).
These resistors were plugged into the inverting input of a 741 op-amp which was powered by a
15 V direct current (DC) with a voltage follower. The equivalent resistance and output voltage
was then calculated over the range: 0˚C to 50˚C. By placing the resistors in this configuration,
the equivalent resistance will be equal to the thermistor’s 37˚C resistance at low temperatures
and equal to RT at high temperatures.

Figure 2

Direct flow rate

The direct flow rate was measured by measuring the amount of time it took for an aquarium
water pump to eject 100 mL of water. The time was measured with a digital timer.

Thermodilution

The thermodilution flow rate was calculated using the Stewart-Hamilton equation after diluting
the pump water by 2˚C water. The 2˚C water was injected into the flow of the aquarium pump
using a syringe. The syringe was connected to the aquarium pump using a y-connector. The
temperature of the pump-cold water system was measured downstream of the injection point
using the thermistor we created. Additionally, the thermodilution rate was determined at a fast,
half fast, quarter fast injection rate. Then, the fast injection rate was used to measure a pinched
dilution rate. The pinched flow rate was induced using a c-clamp placed upstream of the output
flow.

Results

Thermistor calibration curve


12000

10000
Resistance (Ω)

8000
Temp RT
6000
25 11170.21
4000 30 8700
35 6857.143
2000 37 6168.224
0
40 5504.587
0 20 40 60 80 50 2521.008
Temperature (˚C) 60 1983.471
70 1463.415

Linearization of Thermistor

70

60
y = -23.52x - 41.19
50
Temperature (˚C)

Temp Vi 40
27 -2.8
5 -2 30
10 -2.2 20
20 -2.6
30 -3 10
40 -3.4 0
50 -3.8 -5 -4 -3 -2 -1 0
60 -4.4 Voltage (V)
63 -5.2
Direct measurement of flow rate

Volume (mL) Time (secs) Flowrate (mL/secs)


100 25.31 3.951007507
100 30.98 3.227888961
100 29.57 3.381805884
3.520234117
0.380915857

Thermodilution plots
Calculated flow rate

Flow rate (mL/secs)


Fast 2.59
Half fast rate 1.75
Quarter fast rate 0.88
Pinched 1 2.14
Pinched 2 1.48

Discussion

The purpose of this lab was to determine flowrate using the thermodilution method and compare
it to direct flowrate measurements. The thermodilution method required the injection of a small
volume of ice cold water to cool down the flowing water. The change in temperature from
baseline levels was measured using a thermistor and the Stewart-Hamilton equation was used to
calculate the flowrate.

The average (standard deviation) direct flowrate of the flowing water was 3.52 (0.89) mL/secs.
The flow rate calculated using the thermodilution method was 2.59 mL/secs, 1.75 mL/secs, and
0.88 mL/secs when a fast, half of the fast and a quarter of the fast ice water injection speeds were
used. Based on the results, the accuracy of the flowrate is directly related to the injection rate.
This result is due to the time it takes for the temperature of the flowing water to reach the
baseline temperature from the drop in temperature. Additionally, the flowrate of the flowing
water was calculated when the flowing water was pinched and the fastest injection speed was
used. The average (standard deviation) of the pinched rate was 1.81(0.47) mL/secs.

Through this lab, we were able to combine instrumentation and analytical tools to develop a
method of determining flowrate in vivo. The lessons learned in this lab can be applied to other
projects in which it is necessary to measure flowrate.
Appendix I
%% Importing data

data = xlsread('Pinched 2.xlsx');


temperature = data(2:end,2);
time = 0:1/100:(length(temperature)-1)/100;

% convert samples to seconds


%% Voltage to Temperature

temperature = (-23.52*temperature)-41.19;

temperature = tsmovavg(temperature,'s',100,1); %moving average filter with a


lag of 100
plot(time,temperature)
xlabel('Time (seconds)');
ylabel('Temperature (degree Celcius)');
title ('Pinched 2 Injection rate')

%% Flow rate

V_i = 100; %mL


int_T = trapz(time(100:end),temperature(100:end));
T_b = 2;
T_i = nanmean(temperature(1:100)); %average of the baseline temperature.
Change depending on the case

%Steward-Hamilton
F_b = ((T_i - T_b)*V_i)/int_T;

Вам также может понравиться