Loading...
Loading...
Loading...
Loading...
Loading...
Loading...
Loading...
Loading...
Loading...
Loading...
Loading...
Loading...
Loading...
Welcome to AE3610 - Experiments in Fluid and Solid Mechanics
Here you will find the lab manuals for all experiments that will take place this semester. Navigate to the appropriate experiment on the left-hand menu. You can export each experiment as a PDF if you want a copy for your records, or to print as a hard copy. Also, bookmark this page on your smartphone or tablet as well as your PC/laptop as a handy guide during your lab session.
Note that this semester is the first-time we are rolling out Gitbook to deliver the manuals, so please let your TA or Instructor know if there are any typos, or suggested improvements...
Best of luck for the semester!
Excel is great for one-off activities but repetitive tasks are not easy to execute, and large data sets cause Excel to slow down heavily. This is where MATLAB comes into its own... Embrace it!!!
One of the quickest ways to get data into MATLAB is to simply copy and paste it in from your external source. We most often save data in text (.txt) files or spreadsheet files (.csv, .xls, .xlsx), so we consider manual import methods for those two types of file here.
Consider the data obtained from a drone motor electronic speed controller (ESC) opened in Notepad as shown above (it has 505 seconds of data in total). Since Notepad doesn't allow you to select columns, and indeed doesn't even know the data it contains is actually arranged into columns, we must copy it all into MATLAB in one go. Follow the steps as follows:
Since MATLAB doesn't allow a mixture of text and numbers in arrays, first temporarily delete the header row in Notepad
Copy all data in the file by selecting all then copying(Ctrl+A
then Ctrl+C)
Select the top right cell and paste in your data using Ctrl+C.
You don't have to save the variable, it does so automatically.
Return to Notepad and undo your changes to the header or close the file without saving so you don't accidentally modify it forever.
Unless you want to keep the full ESC
array variable for any reason, delete it by right clicking on it in the Workspace and hitting Delete
.
The process for spreadsheets is very similar, but since spreadsheet software allows you to select columns, you can repeat Steps 3-5 above for each column and omit the array column naming in Step 7.
MATLAB also has a powerful Import Data tool which gives you a lot more options for how to import the table, whether to include headers or not, etc. This can certainly be used, too, but has a few more options to know about compared to the Manual method. If you do go this route, to get the same kind of data type as above, use 'Numeric Matrix' as the Output Type. An advanced, but highly powerful, Output Type is the 'table' which has a lot of benefits that are not needed for this simple data reduction.
It is tempting to write your code in the command line, but this can quickly lead to a mess of variables and forgotten discoveries. Furthermore, it is very difficult to execute loops and other multi-line statements. Instead, start a script and modify that as you go a long, thus keeping an ongoing record of what you've need.
A really cool feature in MATLAB is the 'Live Script' - this is a script that also enables you to insert and format text, equations, images, etc. Moreover... it can be exported to a PDF so it shows all your code and the output plots without you having to format anything. One final nice touch is that you can set up variables as interactive inputs, so you don't have to re-run the script to change parameters... it just automatically updates.
Sometimes we will continuously gather data over a long time period, when only certain periods are of interest. Take for example the ESC data from the Importing Data for MATLAB section... this was a power consumption test of a drone propeller where only the power during periods of static throttle are of interest, not the transient periods. Below you will find a PDF of a simple Live Script used to perform such a time series reduction. The Live Script and data are also included for you to play with if desired.
Sometimes we won't collect data over time, instead gathering it in a discrete bin or category. If there are a large number of bins it can take a long time to separate and perform calculations on this data. But fear not... MATLAB eats this sort of problem for breakfast thanks to logical indexing and the ability to loop. In this example, a data set containing noisy atmospheric pressure measurements gathered by 50 weather balloons is binned by altitude. Open the Live Script-exported PDF below to show how to deal with this sort of problem. The data and files are also included for you to explore:
The primary objective of this experiment is to familiarize the student with digital data acquisition of time-varying signals. This lab covers concepts in frequency analysis of time varying signals and sampling theory. It also provides an introduction to computer-based data acquisition systems. In this experiment, you will use a computer data acquisition system to sample signals produced by waveforms stored in mp3 formats and converted to analog electrical signals by the computer’s audio system. You will explore issues in sampling, including the Nyquist limit and aliasing, and the use of analog filters. This will help prepare you for future experiments in this laboratory course that employ computerized data acquisition and involve frequency based interpretation of measured data.
Most experimental measurements involve the dimension of time. Experimental data is acquired over the course of some time, and during this time the signal can change. In some cases, the actual physical parameter of interest (the measurand) may change with time. For example, the velocity in a wind tunnel generally varies with time due to turbulence or variations in the speed of the fan blades used to drive the tunnel. Even when the measurand is nominally constant in time, other parameters that influence the measurement may vary, for example drifts in the measurement device. Thus, the experimenter is often interested in measuring a variable that could be described by the general function (or waveform),
(1)
One of the simplest time-dependent functions we encounter is the sine (or cosine*),
*Either function is acceptable, since , i.e., the two functions are identical except for a phase difference of π/2 or 90°, meaning that shifted by one-fourth of a cycle, cosine looks just like sine.
(2)
More general periodic waveforms, which repeat themselves with a period T and thus have a frequency f=1/T, can be written as a linear combination of simple harmonic modes. There is the basic, fundamental mode (with frequency f ), and harmonics of the fundamental mode, with integer multiples of its frequency (2f, 3f, …). For example, we could describe the vibrations of a tuning fork or the acoustic oscillations in a pipe this way. Mathematically, this linear combination of modes is expressed as a Fourier series expansion,
(3)
Figure 1. A waveform composed of a fundamental mode (at 50 Hz) and its 9th harmonic (at 10 times the fundamental frequency, or 500 Hz). The waveform also has a DC, or time-averaged, component of 4 mV. Specifically, the signal (in millivolts) is 4+sin(100πt)+2sin(1000πt), or equivalently, based on cosines, 4+cos(100πt-π/2)+2cos(1000πt-π/2), which simply represents a phase shift of -π/2.
For example, Figure 1 shows a simple waveform composed of two frequencies, a fundamental mode at 50 Hz and its 9th harmonic (at 500 Hz). Thus the complete waveform is repeated every 20 ms (=1/fundamental frequency =1/50 s). The waveform shown in the figure also has a DC component. In other words, the signal has a nonzero value when averaged over its period. In general, we can write the DC amplitude as
(4)
The other coefficients of the Fourier expansion are given by
(5)
(6)
The power, P, contained in single mode is given by the square of the amplitude
(7)
(8)
A second example that shows the ability of a combination of sine waves to create an arbitrary periodic function is shown in Fig. 2. Five sine waves and a DC component (see Fig. 3) were combined to create a function approaching a square wave. While the constructed function resembles a square wave, it is clear that more sine waves would be needed to produce a sharp square wave.
Figure 2. Partial reconstruction of a square wave using five sine waves, each with a different amplitude, frequency and phase, and a separate DC component. The individual waves are shown in Fig. 3.
Figure 3. The five sine waves and constant function used to construct the square wave shown in Fig. 2.
(9)
(10)
For example, Figure 4 graphically shows the Fourier transforms of various functions, including sine and cosine waves, a rectangle function (Π), a triangle function (Λ) and a constant, or DC, function. The sine, cosine and DC waveforms result in Fourier transforms that are nonzero at a single frequency**;
Figure 4. Fourier transforms of various functions (left and right pairs). The arrows represent impulse functions (i.e., delta functions), which extend infinitesimally along the x-axis, but have a integrated area corresponding to the height indicated by the arrow. The dashed regions indicate imaginary values.
(11)
where it is sufficient to consider only 0<f<∞ since the PSD of a real function is symmetric about f=0.**
Extensions of the Fourier Transform method have been developed for non-continuous functions, specifically for signals that have been discretely sampled by a computer, data acquisition system, or produced by digital means. These are generally known as Discrete Fourier Transforms. In addition, methods to quickly compute the Fourier Transform have also been developed, e.g., the Fast Fourier Transform. These concepts are described in detail in references 2 and 4. The computer data acquisition system you will use employs these techniques to compute the power and phase spectra of the signals that are sampled in this lab.
As a simple example, consider a single sine wave. If we know we are dealing with a single frequency sine wave, it takes at least two measurements per period to determine its frequency, which means we must sample at twice the sine wave’s frequency. If we sample any slower, we actually infer a lower frequency than the actual frequency of the sine wave (you will see this in the lab). This process, by which information at a higher frequency shows up at a lower frequency is known as aliasing.
Data will be acquired with an board in a computer, utilizing a LabView™ interface. The computer data acquisition board, which measures the voltage of the input signal, essentially consists of a multiplexer, a sample-and-hold device, an analog-to-digital converter, a memory buffer, an interface to the computer’s memory, a master clock, and controller (see Figure 5).
Figure 5. Schematic of multiplexed, sequential sampling, computer data acquisition board and its connection to the computer.
The multiplexer (MUX) is a switch that connects one of a number of input channels (usually numbered starting at 0) to the sample-and-hold (S/H). The input voltage on the channel switched by the MUX “charges up” the sample-and-hold during some time interval, which is a fraction of the sampling period (the time between samples). This circuit is then disconnected from the input voltage, and some of the stored charge is drained from it. From this charge, the original voltage value connected to the S/H is determined, and the result is converted to a digital value by the analog-to-digital converter (ADC). The digital value (sometimes referred to as a “word” of data depends on the input voltage, the voltage range of the ADC (the minimum and maximum voltages it reads, e.g., 0-5 V), and its digital dynamic range (number of “bits” = N). The relation between the digitizer output and the voltage input is given by
(12)
Multiple signal inputs are recorded by using the MUX to cycle through each of the input channels at a rate that must be faster than the overall sampling rate (how often a given channel is read) times the number of input channels being read. In the sequential sampling system illustrated in Fig. 5 (and which is representative of the system you will be using), note that the channels are not read at exactly the same time. There is a time delay (skew) between when one channel and the next is read. The skew determined by the maximum switching and reading rates of the MUX, S/H and ADC. This is illustrated in Fig. 6. Simultaneous data acquisition systems, which have negligible skew, typically employ multiple, synchronized S/H systems just upstream of the MUX (see Fig. 7).
Figure 6. Time delay (skew) between successive channels in sequential sampling system.
Figure 7. Schematic of simultaneous sampling, computer data acquisition system.
You will also use an analog, electronic filter manufactured by Krohn-Hite. It actually contains two filters, which separately can be switched to be either low pass or high pass filters. The cutoff-frequency of each is also adjustable, using a combination of a dial and multiplier setting. You will be examining the effects of the filters on the following time-dependent signals:
Take a tour of the equipment with the TAs. In particular you will look discuss:
The MP3 player
The Krohn-Hite filter
Our LabView VI
The National Instruments DAQ (data acquisition) device
The Tektronix oscilloscope
Miscellaneous cables and connectors
Prepare the cables/connections:
Using the 3.5mm-to-BNC cable, connect the MP3 player's audio output to the BNC T-connector
Plug the T-connector into channel AI0 (Analog Input 0) on the DAQ with the toggle set to FS (floating source)
Using a coaxial jumper cable, connect the T-connector to Channel 1 on the oscilloscope.
Note: cables can go bad, especially BNC connectors. If you experience signal dropouts or unexpected noise, have the TA check your cables.
Prepare the LabView VI:
Open DigitalSampling.VI if not already open. It is located in D:\AE3610\<SemesterYear>\1. Digital Sampling\Execution Files
Hit the Run button
Set the sampling rate to 22,000 Samples per second, the number of samples recorded to 6000, and the averaging level to 1.
Prepare the MP3 player:
Power on the MP3 player and ensure it is not connected to power (this will introduce noise into the audio output)
Set the output (volume) level to 32
Enable track repeat
Prepare the oscilloscope:
Power it on using the push button on the lower left
With the help of the TAs, ensure that all key switches/dials are in the correct position
Commence waveform identification and characterization:
There are 11 audio tracks loaded onto both the MP3 player and the PC's hard drive. Each audio track contains a different periodic signal. These signals include: single sine waves, a sum of three sine waves, a product of two sine waves, and periodic waveforms that are not sine waves: square waves, triangle waves, and ramps. Some tracks also have “noisy” signals.
Beware: for compounded signals, power spectrum frequencies only decompose the waveform as if it is a sum of sines!!!
As you analyze each track, simultaneously pull up the same track on the PC and play it through the speakers using media player software, for auditory context.
View the output of each track on the oscilloscope, the VI time plot, and the VI power spectrum, having adjusted their display parameters to best see the waveforms. For each track, take notes/data that best describe the waveforms. These notes should include waveform shape/description, and the approximate frequency of any/all peaks in the power spectrum. If you wish, you could also take screenshots/photos of each display for future reference.
Tip: With the waveform displayed as you like, toggle the Continuous/Hold switch to the Hold position so that it is "frozen" in the frame. Drag the markers in the power spectrum to help you identify specific numbers.
Gather data to understand Nyquist sampling theory and aliasing:
In this step we will determine how aliasing, brought on by sampling rates below the Nyquist frequency, affects our ability to accurately reconstruct/analyze a signal.
Disconnect the T-connector and connect the 3.5mm-to-BNC cable directly to the DAQ's AI0 port. Turn off the oscilloscope.
Find and play the track containing the 1 kHz sine wave.
With a sampling rate of 2500 S/s and record length of 2500 S, acquire a power spectrum. Record at what frequency in the spectrum the peak occurs (i.e. the frequency with the maximum power).
Repeat the above step for the following seven sampling rates: 2000, 1500, 1200, 1000, 800, 675 and 665 S/s, in each case setting the record length value to the sampling rate value (i.e. capture X samples at X S/s).
From the 8 observed frequencies, identify at which sampling rate(s) aliasing is occurring.
For at least two additional sampling rates of your choice (below 650 S/s, with a matching record length as before), first predict whether aliasing will occur. If you believe aliasing will occur, predict the specific aliasing frequency, then acquire data to verify experimentally.
Gather data to explore the effects of varying record length and sampling rate:
Some important terms:
Sample = a single measurement captured by the DAQ
Record = a batch of samples collected by the DAQ before downloading to the VI
Record length = the number of samples in a record
Record time = the period over which the record was captured
Sampling time = the period between successive samples
Sampling rate = the number of samples acquired in a given period of time
Frequency resolution = the frequency spacing between two points in the power spectrum
Again play the track containing the 1 kHz sine wave.
Explore the implementation of a low pass filter to remove unwanted noise:
Reconfigure the cables/connectors:
Remove the BNC cable from AI0 and replace the T-connector back into AI0.
Connect the 3.5mm-to-BNC cable from the MP3 player to the T-connector.
Connect the remaining open port on the T-connector to either input of the Krohn-Hite filter using a coaxial jumper cable.
Connect the corresponding output from the Krohn-Hite filter to AI1 on the DAQ, ensuring the switch is set to FS (floating source), with a coaxial jumper cable.
Power on the filter and set it to LOW PASS x100 mode
Locate and play the track containing the sum of three sine waves at three frequencies on the MP3 player.
Set the following VI parameters:
Sampling rate = 22 kS/s
Record length = 1000 S
Number of averages = 1
Display Settings
Window = None (Uniform) | Vrms | Linear
Plot = Amplitude | Radians
In both time plots, turn off x-axis autoscale and set the limits from 0 to 0.005 s (will round up to 0.1 after hitting enter)
In this step, the sum of sines represents a fictional scenario whereby a signal with two low frequency components of interest (e.g. vibration data from a structures experiment) are subject to a high frequency noise component. Your goal is to remove the high frequency noise without altering the two low frequency components of the signal of interest. This is a very common scenario for signal processing in engineering and science.
Paying attention to the time history and power spectrum of both filtered and unfiltered signals, adjust the cutoff frequency dial of the low pass filter until you obtain a cleaned up filtered signal. When you are happy with your results, take a screenshot of the VI for your report and note down the cutoff frequency on the filter.
Explore the ramification of low pass filtering on signal phase
Continue playing the previous track (sum of three sine waves).
Set the Krohn-Hite filter to a cutoff frequency of 10 kHz
Set the following VI settings:
Sampling rate = 22 kS/s
Record length = 1000 S
Number of averages = 1
Display Settings
Window = None (Uniform) | Vrms | dB
Plot = Phase | Radians
Set the VI to CONTINUOUS and press HOLD to freeze the plots after a few seconds, once new records have downloaded.
Successively zoom in to the 3 frequencies of interest in each phase plot by adjusting the x-axis limits.
Record the phase of the unfiltered and filtered signals at each frequency. Have the TAs check your data.
Gather data to determine the transfer function of the low pass filter at 800 Hz
Locate and play the repetitive sweeping track on the MP3 player
Set the following VI settings:
Sampling rate = 6000 S/s
Record length = 3000 S
Number of averages = 1
Display Settings
Window = None (Uniform) | Vrms | Linear
Plot = Amplitude | Radians
Set the Krohn-Hite cutoff frequency dial to 800 Hz.
With data acquiring, click HOLD and then TAKE NEXT as many times as needed to display a relatively flat and noise-free power spectrum. The TAs will help you achieve this.
When you are happy, click SAVE and choose a useful filename, being sure to add .xls as an extension (this can be added in Windows Explorer afterwards if this step is forgotten).
Change Number of averages to 10, repeating the previous 2 steps to acquire a new data set.
Explore the implementation of a band pass filter to remove all frequency content except for one frequency of interest:
Locate and play the excessively noisy single sine wave track on the MP3 player. Ensure that the MP3 player volume is set to 32.
Set the following VI settings:
Sampling rate = 22000 S/s
Record length = 5000 S
Number of averages = 1
Display Settings
Window = None (Uniform) | Vrms | dB
Plot = Amplitude | Radians
Remove the output of the lowpass filter from DAQ AI1 and instead connect it to input of the currently unused channel of Krohn-Hite filter.
Connect the output of this channel to DAQ AI1 using a further coaxial jumper cable.
Set the second filter channel to HIGH PASS x1 mode and set the cutoff frequency dial such that the frequency is as low as possible (around 20 Hz).
Change the original low pass filter to x100 and set the cutoff frequency dial such that the frequency is much higher than the frequency we are trying to preserve (around 20 kHz).
On both of the filtered and unfiltered time history plots:
Turn off x-axis auto-scale and set the x-axis limits between 0 and 0.01 s.
Turn off y-axis auto-scale and set the y-axis limits between -0.1 and 0.1 V.
On both power spectrum plots:
Turn off x-axis auto-scale and set the x-axis limits between 0 and 11 kHz.
Turn off y-axis auto-scale and set the y-axis limits between -130 and -30 dBVrms (-30 on the top-most limit).
Take a screenshot of the VI for future reference.
Gain a high-level perspective of how both filters affect the power spectrum:
Whilst watching the filtered power spectrum and the time history plots, slowly reduce the low-pass filter cutoff frequency to 2000 Hz. Observe how both plots change.
Zoom closer into the frequency of interest by adjusting the x-axis limits of both power spectra to between 0 and 2000 Hz.
Whilst watching the filtered power spectrum, slowly increase the high-pass filter cutoff frequency to 200 Hz. Observe how the power spectrum and the time history plots change.
On each power spectrum, drag marker "m1" until it snaps onto the peak of the frequency of interest. The y-axis readout (in orange below the x-axis) for each marker tells you the dBVrms of each marker, filtered and unfiltered, and thus the magnitude of the signal at that frequency. If the above steps have been followed correctly, you should see almost identical dB values at these points; have the TA check your setup if this is not the case.
Note down the dBVrms of each peak.
Dial in the cut-off frequencies to complete your band-pass filter design:
Keeping the low pass cutoff frequency at 2000 Hz, adjust LOW PASS mode to x10.
Keeping the high pass cutoff frequency at 200 Hz, adjust HIGH PASS mode to x10.
Successively adjusting both dials until the filtered power spectrum has as much unwanted frequency content removed as possible, without exceeding a 3 dBVrms drop at the frequency of interest. The final filter cutoff frequencies should be evenly spaced around the frequency of interest (i.e. 1000 +/- X Hz).
Note down your final band-pass filter frequencies, the final dBVrms values of each peak (unfiltered and filtered).
Take a screenshot of the LabView VI at its current zoom level.
Set both power spectrum x-axis limits between 0 and 11 kHz before taking another screenshot of the LabView VI.
Lab shutdown procedure
Plug the MP3 player in to charge
Unplug all coaxial cables and arrange neatly on the desk
Turn off filter, oscilloscope, and DAQ
Have the TA upload your data/files to Canvas
The waveform shape/description and peak frequency(s) for each of the 11 tracks.
For the 1 kHz waveform:
For the 1 kHz waveform at the 9 different sampling rate/record length combinations:
For the sum of 3 sines waveform:
For the sum of 3 sines waveform run through a lowpass filter set at 10 kHz:
For the sweeping sine waveform run through a lowpass filter set at 800 Hz:
For the excessively noise sine waveform:
From your data taken in step 3 of the Data to be Taken section, come up with rules for:
The number of points in the power spectrum as a function of record length
The power spectrum frequency resolution as a function of record time
The power spectrum frequency resolution as a function of sampling rate and record length
The power spectrum x-axis limit as a function of the sampling rate.
From your data taken in step 5 of the Data to be Taken section, calculate the phase difference between the unfiltered and filtered signals at all 3 frequencies of interest.
Calculate the ratio of the power spectra for the filtered data (i.e., the output from the filter) and unfiltered data (i.e., the input to the filter) recorded in step 6 of the Data to be Taken section. Find the Transfer Function of the filter as a function of frequency. Do this for both the single record run and the average of 10 records run.
Note: This is a Data Report, so you must follow instructions on how to prepare the Data Report (not the FORMAL report) on the Canvas course page. Also, be sure to answer any supplemental questions listed on Canvas for this lab.
A table describing of each of the 11 tracks including all relevant characteristics (Step 1 of Data to be Taken).
A table containing the data from Step 2 of the Data to be Taken section.
A table containing the data from Step 3 of the Data to be Taken section.
All 4 equations for Step 1 of the Data Reduction section.
A table containing the design cutoff frequency of the lowpass filter (step 4 of Data to be Taken section).
A figure of the lowpass filtered and unfiltered time history and power spectra from step 4 of the Data to be Taken section.
A table containing the data from step 5 of the Data to be Taken Section and Step 2 of the Data Reduction section.
Plots of the single and average power spectra, filtered and unfiltered, for the waveform with the rapidly sweeping frequency (step 6 of the Data to be Taken section)
Plots of the filter transfer function described in step 3 of the Data Reduction, one based on 1 record and another based on the 10-record average results.
For the bandpass filter design (Step 7 of Data to be Taken):
A table containing:
Three screenshots of the LabView VI during filter development
R. V. Churchill and J. W. Brown, Fourier Series and Boundary Value Problems, 3rd ed., McGraw-Hill, 1978.
R. N. Bracewell, The Fourier Transform and Its Applications, 2nd ed., McGraw-Hill, 1978.
T. G. Beckwith, R. D. Marangoni and J. H. Lienhard V, Mechanical Measurements, 5th ed., Addison-Wesley, 1995.
W. H. Press, S. A. Teukolsky, W. T. Vetterling and B. P. Flanner, Numerical Recipes - The Art of Scientific Computing, 2nd ed., Cambridge University Press, 1992.
The primary objective of this experiment is to familiarize the student with the measurement of static and stagnation pressures, and (indirectly) velocity, in a subsonic wind tunnel. Static taps and stagnation (Pitot) probes will be used to measure pressures on the surface of a 2D airfoil, in the wake region behind the airfoil and in a boundary layer next to the wind tunnel wall.
When a 2D airfoil is placed in a uniform subsonic freestream, the flow velocity near the airfoil is modified and, as evidenced by the Bernoulli equation, so is the local static pressure. The resulting chordwise pressure distribution on the surface of the airfoil may be calculated by various methods using an inviscid fluid model.* At moderate angles of attack, the flow accelerates over the upper surface of the airfoil, the surface static pressure is less than freestream over most of the chord, and the pressure coefficient, which is defined as
*From AE 3030 and Chapter 4 (Fig. 4.25) in Andersen’s Fundamentals of Aerodynamics
along the airfoil upper surface has mostly negative values. Normally, there is a large suction peak (large negative value of ) very near the leading edge on the upper surface, followed by a region of increasing static pressure (adverse pressure gradient) from there to the trailing edge. On the lower surface of the airfoil, there is a stagnation point near the leading edge, where = 1.0, and the flow accelerates thereafter. When the two pressure coefficient distributions are plotted versus chordwise location, (x/c), the area between the two curves is a measure of the normal force coefficient on the airfoil and hence of the airfoil lift coefficient.
As the angle of attack is increased, the suction peak on the upper surface grows larger and the adverse pressure gradient becomes larger as well. At some value of angle of attack, the adverse pressure gradient on the airfoil upper surface becomes strong enough that the boundary layer separates from that surface. At a sufficiently high angle of attack, the oncoming freestream flow perceives a radically modified airfoil shape. The resulting effect is termed airfoil (or wing) stall, and the included area between the upper and lower pressure distribution curves collapses. The presence of stall was evident in the force measurements on the airfoil that were conducted in a previous laboratory.
The airfoil used in this experiment (see Figure 1) has an NACA 64212 section. The chord is 14 inches and it has a thickness ratio of 14%. The airfoil extends from one side wall of the wind tunnel to the other. Therefore, it should behave like a 2D airfoil.
Figure 1. Two-dimensional wing showing pressure tap locations (light colored line at centerline of airfoil), red tufts used to visualize separation and tubing used to connect pressure taps to Scanivalve.
In the flow of a viscous fluid such as air, the flow velocity right at a solid surface is zero; i.e., the fluid can be thought of as adhering to the surface (the no slip condition). Within a small interval above the surface, the flow velocity increases rapidly from zero to a value that is of the order of the freestream velocity. The result is a velocity profile which exhibits a large velocity gradient in the direction normal to the surface. This velocity gradient gives rise to significant shear stresses, and the region within which this takes place is termed a boundary layer. Using boundary layer theory** one may show that the static pressure is constant through the boundary layer in a direction normal to the surface and that the boundary layer is a region of rotational flow so that the stagnation pressure is not constant everywhere. However, the Bernoulli equation may be used locally to find the dynamic pressure distribution within the boundary layer and hence the velocity profile if the flow is incompressible.
**See your textbook from AE 2010 and 3030. It would be helpful if you review this material before coming to the lab if you do not know what the velocity profile in the boundary layer looks like.
In the case of viscous flow over an airfoil at a moderate angle of attack, the attached boundary layers on the upper and lower surfaces join at the trailing edge. The resulting viscous-dominated flow region downstream of the trailing edge is termed a wake, within which there is a velocity deficit compared to the freestream. This deficit is a result of the flow being retarded in the airfoil boundary layers.
The surface pressure distribution on an airfoil will be measured by means of 24 static pressure taps. These are small holes on the surface of the model that are connected to stainless steel tubes within the model and thence to plastic tubing, which are ultimately connected to a pressure transducer. The pressure taps on the airfoil are located on the upper and lower surfaces in the chordwise direction at mid-span, according to the following chord-normalized horizontal positions:
Table 1: "Tap map" showing the chord-normalized positions of each airfoil static tap
The local velocity in the boundary layer on the ceiling of the wind tunnel and in the wake behind the airfoil will be measured indirectly by traversing a Pitot* tube through the region so as to measure local stagnation pressure; the velocity follows directly because the flow is incompressible. Since the static pressure is constant throughout the ceiling boundary layer, a single static tap on the ceiling (ideally at the measurement station) will yield the local static pressure anywhere in the boundary layer at that station. In the wake region, the local static pressure will be approximately constant through the wake and will be equal to the freestream static pressure. This is because the wake measurement station is located sufficiently far downstream of the airfoil that the pressure disturbance due to the airfoil is negligible.
*Named after its inventor, Henri Pitot (1695-1771), a French hydraulic engineer.
The end of the Pitot probe is made of thin stainless steel tubing with a 0.063 inch outer diameter (OD) and a small orifice, so that the Pitot pressure data is a local measurement compared to most other dimensions. Generally Pitot probes such as ours may be oriented a few degrees (say 5-10 degrees) away from the local flow direction without any appreciable change in the measured pressure, hence a precise alignment of the probe with the local flow direction is not required. A Pitot-static probe (a Pitot tube with a downstream static pressure tap oriented normal to the stagnation hole) located upstream of the airfoil will be used to measure freestream dynamic pressure.
We can categorize this type of pressure transducer by the way the deformation of the structural element is transformed into an electrical signal. The two most common approaches are strain gage and capacitance type transducers. A strain gage pressure transducer consists of a thin circular diaphragm on the bottom of which are bonded tiny strain gages wired as a Wheatstone bridge. When the diaphragm experiences a pressure on its exposed upper surface that is different from the pressure in a small cavity under the diaphragm it deflects, and the resulting bridge imbalance is a measure of the deflection. Calibration provides the constant of proportionality between bridge imbalance (interpreted as a voltage) and applied pressure. The voltage output of this type of transducer is usually in the millivolt (mV) range and requires amplification prior to measurement. Relatively inexpensive transducers can be made by using semiconductor materials. In this case, the semiconductor resistors are “written” as a bridge circuit directly onto a substrate (e.g., silicon) that acts as the diaphragm. The strain on the semiconductor results in a change in semiconductor resistance; this is known as the piezoresistive effect. The change in semiconductor resistance is analogous to the change in metal resistors, except in for metal resistors, the change in resistance is primarily due to the change in the resistor’s cross-sectional area as it is strained. For semiconductor materials, the resistance change is related to other changes in the internal structure of the semiconductor.
The capacitance-based pressure transducer has a stretched membrane clamped between two insulating discs, which also support capacitive electrodes. A difference in pressure across the diaphragm causes it to deflect, increasing one capacitor and decreasing the other. These capacitors are connected to an electrical, alternating-current (AC) bridge circuit, producing a high level of voltage output (usually 10 Volts full scale without amplification).
Strain gage transducers can be made small, hence they can be internally mounted in a wind tunnel model. Also, they have reasonably good frequency response because of the small mass of the diaphragm and the short distance between the pressure tap and the diaphragm face. Capacitance transducers usually are not well suited for internal mounting (too large) and such systems do not have a fast response. Both types of transducers can be calibrated using a primary pressure standard such as a dead-weight tester, which supplies a pressure of precisely known magnitude, or using another (already) calibrated pressure transducer. In this lab, you will use both capacitance type and semi-conductor strain gage transducers to measure pressure.
Whereas all previous measuring devices measure only wind tunnel and/or test article pressures, the FlowKinetics 3DP1A device also measures local ambient pressure and temperature, enabling air density and thus flow speed to be calculated. As shown in Figure 5, the FlowKinetics has a selector dial for choosing the display of either pressure or velocity (with multiple options for units on each). The FlowKinetics is an independent reference in that it does not output any signals that are read during the lab; we will be using it only for a sanity check on flow speed magnitude and for the ambient pressure and temperature measurements.
A wind tunnel is a duct or pipe through which air is drawn or blown.* The Wright brothers designed and built a wind tunnel in 1901. The basic principle upon which the wind tunnel is based is that the forces on an airplane moving through air at a particular speed are the same as the forces on a fixed airplane with air moving past it at the same speed. Of course, the model in the wind tunnel is usually smaller than (but geometrically similar to) the full size device, so that it is necessary to know and apply the scaling laws in order to interpret the wind tunnel data in terms of a full scale vehicle. The wind tunnel used in these experiments is of the open-return type (Figure 6). Air is drawn from the room into a large settling chamber (1) fitted with a honeycomb and several screens. The honeycomb is there to remove swirl imparted to the air by the fan. The screens break down large eddies in the flow and smooth the flow before it enters the test section. Following the settling chamber, the air accelerates through a contraction cone (2) where the area reduces (continuity requires that the velocity increase). The test (working) section (3) is of constant area (42" × 40"). The test section is fitted with one movable side wall so that small adjustments may be made to the area in order to account for boundary layer growth, thus keeping the streamwise velocity and static pressure distributions constant. The air exhausts into the room and recirculates. The maximum velocity of this wind tunnel is ~35 mph, and the turbulent fluctuations in the freestream are typically less than 0.5% of the freestream velocity. Thus, it is termed a “low turbulence wind tunnel.”
*Anderson, p. 123
This lab is broken down into 2 major parts:
Surface pressure distribution measurement
Wind tunnel boundary layer pitot traverse survey
In all experiments during this lab, the wind tunnel should be set to maximum "throttle", which should result in a wind speed of around 16 m/s (about 35.5 mph). This may vary based on the day
The TAs will first give you a brief tour of the low turbulence wind tunnel and the associated equipment, paying particular attention to that outlined in the Background section above. Once complete, you will need to perform the following items:
Make a manual note of the ambient pressure, temperature, and air density displayed by the FlowKinetics manometer.
Document the mapping between Scanivalve port number and pressure tap label. Note down which Wing Pressure Port (U__ and L__) is connected to which scanivalve pressure port number (this is vital information for your data reports). Check with the TAs for accuracy.
In this first part of the lab, the primary goals are: (1) to get an intuitive feel of what is happening with the airfoil with the wind on at a variety of angles of attack, and (2) to directly measure the pressure distribution over the airfoil.
Configure the Scanivalve software
Go to D:\AE3610\<Semester Year>\Subsonic\ScanivalveDSALink (Real Time Scanner View)
and open Instance 1 of the Scanivalve executable, setting the IP address to 192.168.1.72
(the top scanner)
Go to D:\AE3610\<Semester Year>\Subsonic\ScanivalveDSALink (Real Time Scanner View)
and open Instance 2 of the Scanivalve executable, setting the IP address to 192.168.1.73
(the bottom scanner).
If either scanner won't connect, make sure all cables are properly seated and the network switch connecting the scanners is turned on
For each software instance, make sure the Scan Settings
are set as per the list below (for now we will have averaging of 1 to have the fastest response with no smoothing to hide anything)
Period: 1000 microseconds
Average: 1
Frames Per Scan: 1000000 (arbitrarily high to prevent scan ending too quickly)
Conversion Units: mmHg
Units: Engineering Units
Time Stamp: No Time Stamp
Zero Correct: ON
SCAN trigger: Internal
Click CALZ
in each software instance to zero-calibrate each scanner
Click Scan
to start the scan, ensuring the sensors do indeed look zeroed
Following the wind tunnel operating procedure and with the help of the TAs, power on the wind tunnel to full speed
Take turns varying the airfoil's angle of attack, paying particular attention to the scanner readings throughout, as well as how the airfoil feels in your grip across its range of motion. Be sure to look for, and beyond, the stall angle of attack
Once everyone has finished, have everyone return to the control room and turn off the tunnel
Now having some insight into the angles of interest, decide amongst your group a minimum of six angles of interest at which you will log data. They must include:
Negative angle of attack (AoA)
Zero degrees AoA
Low AoA below critical angle
High AoA below critical angle
Critical AoA: the angle of attack at which lift is greatest
One Angle in the Stall Region: past critical AoA
Make sure the airfoil is at zero degrees angle of attack
For each software instance:
Stop the scan
Change the Averaging level in Scan Settings
to 100 (this will give some low-pass filtering and smaller log file sizes at the expense of speed of response)
Gather pressure data at your six angles of choice:
!!! Thoroughly read all the sub-steps in this step so you understand what is about to happen !!!
Select one member of your group to be the designated "Angle of Attack Adjuster" for the wing
Hit CALZ
Click Start Data File
, navigate to your group's save folder, and come up with a meaningful name for each scanner so you remember for the data reduction which was which
Create your groups save folder in the following location:
D:\AE3610Data\<Semester Year>\<Section>
Once you are comfortable with the sequence, click Scan
to begin recording data to the log file and wait 10 seconds to gather some data at wind-off condition
Power on the wind tunnel to maximum speed
Once the wind tunnel is up to speed, wait around 10 seconds for the zero angle of attack condition to log some data
Cycle through your target angles of attack. For each target AoA, make sure the actual angle is recorded from the markings on the Angle Finder (it may differ slightly from the planned angle you selected), and that the adjuster doesn't disturb the airfoil for around 10 seconds to ensure good data
Once all target AoAs have been visited, return the wing to 0 degrees AoA. Wait 10 seconds, and turn off the tunnel. Wait for the fan to come to a complete stop
Click Stop
and Close Data File
on both Scanivalve instances
Ensure the data saved correctly by opening the saved data files from both instances. Plot at least one of the valid pressure tap columns*. You should see a trajectory starting and ending in zero mmHg, with a ramp up/down as the fan accelerates/decelerates, and multiple plateaus representing the points at which you stopped to log. If the data hasn't saved correctly, or the plateaus are not pronounced enough, re-attempt this procedure from the beginning. You may have a TA help ensure your data is acceptable
Rename your data files to include the actual angles of attack at which you took data
Close both Scanivalve Instances
*Each spreadsheet output by the Scanivalve software has 33 columns:
Column 1 = Data Index
Columns 2-17 = Pressure at each of the DSA's ports (whatever unit you used during Scanivalve logging - should have been mmHg)
Columns 18-33 = Temperature at each of the DSA's ports (degC). Note that all scanner port pressures are recorded, so you should ignore any ports that are not used by this lab.
To control the traverse and acquire data, we use a LabView VI, shown below in Figure 7. The main display on the VI is the large graph in the center, which shows airspeed derived from the Baratron.
Open the SubsonicWindTunnel Application VI found in D:\AE3610\<Semester Year>\SubsonicWindTunnel
Choose the correct file location D:\AE3610Data\<Semester Year>\<Section>
and give it an arbitrary filename (this file will not be saved).
Click "END TEST"
Under Traverse Position
in the VI, select the VISA Resource Name dropdown and Select COM4. If COM4 is not available, ensure the peripherals are plugged in. Select the Visa Resource Name dropdown and click Refresh. Then select the dropdown again and select COM4.
Start the VI again by pressing the Run button (white arrow on the top left of the window) and choose the correct file location D:\AE3610Data\<Semester Year>\<Section>
and give it a sensible filename (this file will be saved).
Input ambient pressure and temperature from FlowKinetics into the Air Density Calculation
part of the VI, which should yield a realistic air density (note that the graph will not input until this is done, otherwise it will be trying (unsuccessfully) to plot NaNs)
Zero the Baratron voltage by clicking Tare on the VI. Make sure you do this with the wind tunnel off, and check that the airspeed reading is centered around 0 m/s. If not, hit Tare again.
Take a screenshot of the VI (like below, but with values populated) to remember your settings and have the image as a reference for future use.
If the traverse ever hits a physical limit, always immediately hit the E-Stop paddle to prevent damage and/or injury. The stepper cannot move without power, even if the Arduino continues to send pulses to it.
If the traverse hits a limit it will typically make a grinding noise so keep an ear out for that happening. You will also possibly see the Arduino Motion Status reporting something different from what you see. In either case, initiate the E-Stop procedure and re-initialize.
Since the Arduino and stepper motor are not synchronized (that is to say there is no feedback of the actual stepper position, only a running tally of the number of pulses sent to it which translates to linear position), it is crucial to initially synchronize the two. To do this, the Arduino commands the stepper downwards until it hits the bottom limit switch, at which point it moves slightly up to de-activate the switch and sets its current position to home.
Initial power on and zero:
Check the position of the traverse carriage. If it is clear of the bottom limit switch, no action is necessary.
If it is touching the limit switch, have a TA turn off power to the stepper motor using the E-STOP and manually twist the aluminum shaft coupler until the carriage is clear of the bottom limit switch. Turn the power back on once the carriage is clear by pressing the GREEN E-STOP button.
Have a TA power cycle the Arduino by pressing its reset button (top-most of the two small black buttons on the circuit board). Alternatively, you can unplug it from the USB hub under the wind tunnel and plug it back in.
Press the GREEN E-STOP button if you have not already done so.
Prep the LabView VI for serial communications with the Arduino:
Stop the VI running if it isn't already stopped by clicking "END TEST"
Under Traverse Position
in the VI, select the VISA Resource Name dropdown and click Refresh. Select the correct device in the list (COM4) and run the VI. Ensure you are in your group save folder and select an arbitrary file name. Successful communication is known to have occurred if the Arduino Motion Status
and Arduino Time
start outputting legible values.
As of 8/24/2023 the correct device is COM4.
Note that the VI only polls the VISA Resource Name once on initialization... changing the dropdown mid-experiment will not change anything. To make changes, you must stop the VI, change the dropdown, and then re-initialize the VI.
Occasionally, nothing can be done within the VI to get the serial communication going. In this event, fully shut down LabView (including the shortcut window that comes up when you close the VI) and open up the VI again, whilst simultaneously having a TA power cycle the Arduino.
Click "Go Home" - the traverse will move down slowly until it hits the bottom limit switch, then goes up a couple of millimeters before setting this position as Home.
Re-synchronization procedure - If at any point the Arduino and stepper motor become unsynchronized, repeat the procedure outlined in Step 1 above.
Traverse reset procedure - Since there is no software stop command in the LabView VI, and the traverse sequence takes a good period of time, you can reset the traverse by pressing the Arduino's hardware reset button (top-most button of the two small black buttons on the circuit board) or power cycling it by unplugging and re-plugging the USB. You may also want to restart the VI to get a new, clean, log file.
For the following three tests, if the Labview VI Application starts lagging suddenly and/or if the plotting becomes discontinuous, restart the VI. Do this by ending the test, closing the entire application, and restarting the program. Do this ONLY IF TIME PENDING!!
Prep the probe for the boundary layer survey:
Run the VI and select an arbitrary file name in your group's data save folder
Click on "Go Home"
Have a TA install a plug over the wake stagnation hole if it is not already there
Re-zero the Baratron in the VI
Following the wind tunnel operating procedure, turn on the wind tunnel and set it to maximum speed
Once the wind tunnel is at maximum speed, click Initiate Boundary Layer Traverse
at which point the hard-coded survey sequence will commence as follows:
Return home
Set Motion Status to LOGGING for "X" seconds
Note: The LabView VI will automatically log when it sees the Motion Status LOGGING flag
Traverse "Y" inches to the next logging position, during which time the Motion Status will read RUNNING
Repeat until the final position is logged
Return home
After the probe returns Home, stop the VI and power down the wind tunnel
Open the log file in Notepad, copy and paste its contents into a new MS Excel spreadsheet, and save the file with a meaningful file name into your groups data folder
Make a quick plot to sanity check the data is all there, if not you will need to re-run the experiment. The file format is as follows: Column 1 = Time (s), Column 2 = Traverse Position (inches), Column 3 = Baratron Output (Volts).
Note (As of 1/16/2024): --The boundary layer stagnation probe is positioned just outside the sidewall of the tunnel at the start of the boundary layer survey. It traverses sideways 4 inches during the test.
Once you have verified all data has been successfully acquired:
Follow the wind tunnel operating procedure to power down the tunnel
Unplug both power cables in the control room; peripherals and e-stop
Close the VI without saving
Have your TA upload your data files from the Wind Tunnel PC to Canvas
Make a tubing schematic of all the probes/taps/instruments in the experiment. Peruse the "Results Needed for Report" section below to see what exact information is expected for your schematics. You may run your schematic by the TAs to ensure it is correct.
Be VERY CAREFUL not to pull the tubes out of the ports!!
A tubing schematic showing how the pitot-static probe, the ceiling static tap, and the traverse probe are connected to the Baratron, FlowKinetics, and the Scanivalves
A diagram or table documenting the Scanivalve port numbers and pressure tap labels
Two CSV files, one per Scanivalve instance, containing the surface pressure of the wing at a minimum of 6 different angle of attacks
A screenshot of the LabView VI containing the ambient temperature and pressure conditions you obtained from FlowKinetics
A LabView log file and corresponding excel spreadsheet containing the boundary layer traverse data
Boundary Layer Survey
Convert the recorded Baratron voltage into wind speed (u) using the Baratron pressure scaling factor, the Bernoulli Equation, and the ambient atmospheric conditions recorded in the LabView VI
δ is defined as the height above the wind tunnel floor where the velocity becomes essentially the same as freestream velocity (>99%)
Make a tubing schematic figure showing all connections necessary for the measurement of (a) the airfoil static pressures and (b) the boundary layer velocity profile. Indicate what each tube is connected to at both of its ends and label each tube as to what pressure it contains. Note: you only need one figure since the tubing setup does not change during this experiment.
A screenshot of the LabView VI containing information about the ambient temperature, pressure, and calculated density.
On one figure, plot the pressure distribution at zero angle of attack as a standard and also plot the pressure distribution at the negative AoA, low AoA, and high (before critical angle) AoA for which you recorded data.
On a second figure, plot the pressure distribution at zero angle of attack as a standard and also plot the pressure distribution at the high (before critical angle) AoA, critical AoA, and stall AoA
Plot data points only; do NOT draw lines or smooth curves between them. Furthermore, average the multiple data points recorded for each static port so you are plotting one point for each static port at a given angle of attack.
On all plots, make sure to plot the lower and upper surface data on the same figure! You should have a total of 2 figures.
This experiment will allow you to investigate hoop and axial stress/strain relations for a pressurized thin-walled cylinder. This is an opportunity to examine a system with a biaxial state of stress, as opposed to the primarily uniaxial stress systems of earlier labs. Furthermore, you will be analyzing transient (as well as steady-state) strain readings from a rosette strain gauge in response to a dynamic event. Finally, you will be able to determine the internal pressure within the vessel from your wall stress measurements and cylindrical pressure vessel relations.
Many components of aerospace vehicles have an internal pressure higher than the surrounding pressure. Perhaps the most obvious examples are the fuselages of manned aircraft and the crew-compartments of spacecraft, which maintain internal pressures close to one atmosphere in low pressure environments. In other applications, the internal pressure can be significantly higher than 1 atm. For example, pressurized structures are required for many gas and liquid storage systems, e.g., propellant tanks in spacecraft, and for the propulsion engines used in aircraft and spacecraft.
We can characterize many of these structures as thin-wall pressure vessels, where the wall is essentially a membrane because its thickness is much less than other characteristic dimensions of the vessel, such as its length or radius. In an ideal thin-wall pressure vessel, the stresses are uniform across the thin wall and the bending stresses are negligible over much of the structure. The vast majority of thin-wall pressure vessels are nearly spherical or cylindrical in shape (or composed of combinations of these two geometries). A sphere provides the advantage of minimizing stresses for a given volume and storage pressure compared to other shapes, i.e., it is structurally efficient. While cylinders are not as efficient and have larger stresses at the ends, which require additional structural reinforcement, they often have significant manufacturing and implementation advantages compared to spheres (imagine the drag induced by a spherical aircraft fuselage).
Structures such as vacuum tanks and submarine hulls may also be characterized as thin-walled pressure vessels, through with internal pressures below the external pressure. For those cases, one must consider wall buckling as a failure mode.
Ignoring the ends, we can calculate the hoop stress by considering the top half of the cylinder (a section of this is shown in Figure 2). We can use a free body diagram to get the necessary relationships for a cylinder of length L, radius R and wall thickness t. First, from the definition of stress, we have
(1)
The wall must support the difference in (vertical) forces on the inner and outer sides of the wall due to the pressure difference, so
(2)
Combining (1) and (2), gives us a relation between the hoop stress and the gauge pressure.
(3)
A similar process looking at a cross-sectional cut normal to the longitudinal axis (Figure 3), where the longitudinal force in the wall must balance the pressure difference across the end of the cylinder, gives
(4)
Solving (4) for the longitudinal stress gives
(5)
As you will be measuring strains in our thin-wall vessel, you will need to convert them to stresses. As usual, this relationship is provided by Hooke’s Law for strain, e.g.,
(7)
(8)
(9)
The equipment used in this lab are pictured in Figure 5. The pressure vessel you will investigate is the common soda can. You will mount on the can either a rectangular rosette strain gauge, which measures strain along three directions, or a T-rosette gauge, which measures the strain along two perpendicular directions. You will use the Vishay 7000 acquisition system to record the reading from the strain gauge. Care must be exercised when handling the soda can. Be careful not to indent or drop the can or shake it so as not to increase the internal pressure. Be sure not to excessively strain the lead wires connected to the T-rosette gauges.
Measure the diameter of the can using digital calipers. Zero the calipers with the jaws in the fully closed position. Measure the diameter of the can by laying the calipers over the top of the can with the can standing upright. Carefully close the caliper jaws around the can being sure not to squeeze the can. Iterate between locating the diameter and allowing the jaws to contact the walls of the can. Conduct a total of 3 diameter measurements and record these values.
Measure the thickness of the soda can Aluminum material from the provided cutout sheet. Be careful not to close the calipers too tightly and deform the material plastically. Collect a total of 3 thickness measurements and record these values.
Mount the strain gauge on your soda can.
If you are using a T-rosette gauge, it should be bonded so that one grid is aligned with the longitudinal axis of the can and the other grid is aligned with the hoop direction. It is best to scribe perpendicular lines on the sanded region of the can to facilitate the alignment of the gauge with the axial and radial directions.
Properly connect the strain gauge wires to the strain gauge card in the Vishay 7000.
Configure the Vishay 7000/Strain smart software interface. Be sure to supply the appropriate values for the strain gauge and transverse sensitivity factors - assign the correct values for the appropriate strain gauge grid. Set the data sampling rate to the maximum possible value (2048 samples/second). The fast sampling rate will allow you to capture some data points associated with the early, sharply rising portion of the strain gauge record when the pressure suddenly changes.
The individual strain gauges will already be configured to their respective 1/4-bridge circuits. Zero each 1/4-bridge circuit and perform standard shunt calibrations.
Shake the can carefully, being sure not to accidentally damage the strain gauge, or the lead wires.
WITHOUT OPENING THE CAN, carefully and gently pry up the soda can tab just a bit so that you will be able to easily get your finger underneath when you are ready to crack open the can.
Be sure to note down which strain gauge grid corresponds to each recorded strain gauge signal. This is especially important in the case of the 3-element rectangular rosette gauge if that is the type of gauge you are using in your experiment.
Can diameter and can wall material thickness.
Strain gauge outputs for the transient can opening test.
Displacement rate and the digital sampling rates for the Vishay system
Axial loads and bending strains for loading and unloading runs for short beam-column.
If you used a rectangular rosette for the transient can experiment (and misaligned it as you were supposed to), then you need to transform the measured strains to the principal strains (maximum and minimum) and corresponding principal strain angles using a similar approach to that used in the cantilever beam bending lab. The principal strains should turn out to be the hoop and longitudinal strains.
Determine the rise times for the hoop and longitudinal strain (magnitudes) that occurred as the can was first opened.
Use the value of the Poisson ratio that you found from the material database in the at home exercise and estimate the radial strain in the can wall. Combine with the measured can wall thickness to estimate the change in wall thickness when the can was pressurized.
From you hoop and axial strain versus time profiles, identify the steady state portions of each curve (before opening the can and afterwards where the strain finally levels out to a nearly constant value). Use data from each steady state portion to estimate the overall change in hoop and longitudinal strains experienced by the can.
Table of can dimensions.
Table of gauge factors and transverse sensitivity factors for each of the strain gauge.
Combined plot of measured strains (in microstrain) as function of time (in seconds) for the can-opening experiment.
Combined plot of hoop and axial strains (in microstrain) as a function of time (in seconds). You may truncate your plots in time as required – but be sure to clearly display the transient and steady state of your data.
Table of overall changes in hoop and longitudinal strains and stresses (reported in MPa) from the initial can state to the final state. Provide stress results using both the full (Equations (8)) and simplified (Equations (9)) Hooke’s Law relations.
In the MATLAB Workspace, right click New
and give your variable a meaningful name. In this case I chose ESC for obvious reasons. This will create a new scalar number with 0 value.
Double click on ESC and it will open up in the Variables viewer where you can see its value.
In MATLAB, create new variables for each column in the array in the command line: Note that the first item in the parentheses defines the row of interest, with the second defining the column of interest. Using a colon instead of a number means 'select all'.
Save the data in a .mat file so you don't have to repeat the above steps for later use - do this by right-clicking anywhere in the Workspace, hitting Save, and choosing a meaningful name (I called mine ESC
again).
To load the data in the future you can simply double click on the .mat file in Current Folder
tab, having navigated to that folder first.
Alternatively, you can do this from the Command Line or in a Script by calling the load() function:
where A is the amplitude, is the circular frequency (e.g., rad/s), f is the cyclic frequency (e.g., cycles/s, Hertz or ), and is the phase, which represents the time-shift of the sine-wave from some reference time that defines t=0. Such a function is often denoted as a simple harmonic waveform.
where nf represents the frequency of the mode (n=1 for the fundamental, n=2 for the first harmonic, etc.), represents the steady component of the waveform, and the , __ are the harmonic coefficients (or amplitudes) of each mode. The steady amplitude, , is often called the DC component of the waveform, in reference to classical electrical power systems, which are either Direct Current (steady) or Alternating Current (sinusoidal with a zero average).
and they can be combined into a complex number (since, ),
and the phase (or phase angle) of a mode is given by
The procedure outlined above for periodic functions can be extended to general functions, which are not necessarily periodic, by considering any arbitrary function to be periodic with an infinitely long period. This approach leads to the Fourier Transform. Given a function , its Fourier Transform is a complex function defined by
in parallel to the complex Fourier function of equation (6). The function represents the information given by transformed from the time domain to the frequency domain. The transformation is nearly identical in the reverse direction, with simply a change in the phase (note the sign of the exponent), i.e.,
**The negative frequencies relate to phase information for the sine and cosine and do not actually represent different frequencies, i.e., for real functions , it can be shown that . That means that if you take the absolute value of V, the part of V below 0 frequency looks like a reflection of the part for f>0.
in other words, they contain information at only one frequency (the DC function, which does not change in time, is associated with a frequency of zero). The Fourier transforms of the rectangle and triangle functions result in and functions, where , which contains information at many frequencies, but with multiple frequency “peaks”.
Instead of looking at the Fourier transform, we often are interested in the power spectrum (or power spectral density, PSD) of a waveform. This represents the amount of power or energy in a region between f and f +df. For real (noncomplex) functions , this is given by
Thus the PSD of the rectangle function, as shown in Figure 4, is the square of its Fourier transform, or (also shown in Figure 4).
In most situations, especially for computer-based data acquisition, the continuous function is sampled (i.e., the data is acquired) at evenly spaced, discrete intervals in time, separated by an amount . The sampling frequency (or data acquisition rate) is thus .
For a given sampling rate, we might ask how accurately the discretely acquired data can reproduce the actual waveform being sampled. The answer depends on the frequency content of the waveform and a special frequency, called the Nyquist frequency , which is half the sampling frequency, i.e., . If the waveform contains no components above the Nyquist frequency, then the waveform can be completely determined by the sampled data (assuming no errors in the measurement).*** This is known as the Nyquist/Nyquist-Shannon Sampling Theorem.
***A waveform that has information in only a limited range of frequencies is called bandwidth limited. Due to phase ambiguity, the sampling frequency should actually be more than twice the maximum frequency in the waveform. For example, a sine wave sampled at 0, , , etc. would always have a 0 result and could be confused with a null function.
Aliasing occurs for any sampled waveform having components with frequencies above the sampling system’s Nyquist frequency, i.e., . One way to remove this problem is to filter the data before it is sampled. This can be accomplished by a low pass filter, a filter that only passes frequencies below some cut-off frequency. One would set the cut-off at or below the Nyquist frequency. The high frequency information is thus removed before it can be aliased. In essence, the filter produces a bandwidth limited waveform.
where output has to be an integer value. For example, consider a 2 V input into an ADC with a 0-10 V range, and an 8-bit digitizer (possible values, or digital values of 0-255). The output of the ADC would be a digital value of 51. The digital result is then moved to the buffer memory on the board, and shifted to the computer memory, usually through the computer’s direct memory access (DMA) system.
In this lab, you control the data acquisition process through a software interface called a LabView virtual instrument (VI). The VI creates a display on the computer screen that lets you think of the data acquisition system as a box with “knobs”, “dials”, and other displays. For this experiment, the VI allows you to control parameters such as the minimum and maximum voltages read by the acquisition board, the sampling rate, and the number of samples recorded.
At the sampling rate and record length combinations shown in the table below, determine record time, power spectrum frequency range, power spectrum frequency resolution, and number of points in the power spectrum. Do this by adjusting the x-axis limits on both the time history and the power spectrum as needed, directly observing and noting down each of the required variables.
A pressure transducer is a device that converts a pressure to a quantity that may be readily measured. For example, a traditional U-tube manometer is a pressure transducer, where pressure difference is interpreted as the height of a column of liquid. This is an example of devices known as gravitational transducers. Modern electronic transducers, which convert pressure into a voltage that may be easily measured by means of a digital voltmeter or an analog-to-digital converter (ADC), are typically elastic transducers. The most common types of electronic pressure transducers use the deformation of a diaphragm or similar structural element to sense pressure.
A common pressure transducer for rapidly changing conditions, based on the piezoelectric effect, will be introduced in later labs.
The Baratron, shown in Figure 2, is a capacitance-based transducer (of a specific brand name) to measure freestream dynamic pressure. For this device, the differential pressure ΔP is related to the transducer voltage V by the relation , where R is the responsivity of the transducer, which for our Baratron is 1.016 mmHg/Volt (1 mmHg is essentially equivalent to 1 Torr). Thus, in order to measure ΔP, we need only measure a simple analog voltage which can be done via a multimeter, DAQ device, or similar. At its maximum sensing range of 10 Torr, our Baratron will output roughly 10V, but other sensing ranges are available for purchase.
The Scanivalve Digital Sensor Array is a bank of piezoresistive pressure transducers you will use to simultaneously measure the 24 pressures on the airfoil surface. The Scanivalve system (Model DSA 3217/16PX) is composed of two units (see Figure 4), each having 16 differential pressure transducers with a maximum pressure range of 0.18 psi (or 5” ). The reference pressure for the Scanivalve array comes from a static pressure tap in the wind tunnel ceiling. The average linearity error for the pressure transducers, as determined by the manufacturer, is ±0.03% of full-scale. Each unit has its own analog-to-digital converter and controller that converts the measured voltages into pressures based on a stored set of calibrations. The units are connected to our data acquisition computer via ethernet.
The Pitot probe is clamped onto a traveling nut that moves along a lead screw mounted vertically underneath the wind tunnel test section (see Figure 5). The lead screw is driven by a stepper motor, which is a pulsed direct current (DC) motor capable of shaft rotation in either direction, with a known holding torque and number of steps per revolution. Every time a 5V pulse is received by the stepper motor driver, it commands the stepper to move one step, thus a sequence of pulses can be sent to achieve different types of motion. In our case, an Arduino is used to generate this pulse sequence and is hard-coded to repetitively extend the traverse a set distance before waiting for a period of time to allow data gathering. Since we are commanding rotary motion to achieve linear motion, we need to know the mapping. This is a property of the lead screw called "pitch", expressed in threads (or revolutions) per unit length, and is known from the manufacturer or by directly measuring the threads. Between the number of steps per revolution of the stepper, SPR (steps per revolution), and the pitch of the lead screw, TPD (threads per unit distance), the number of steps, S, required to traverse a desired distance, D, is thus given by . The AE3610 traverse (as of Fall 2021) has a SPR of 400 and a TPD of 5 threads per inch. Thus, to travel a quarter of an inch, we would require the transmission of 500 steps/pulses by the Arduino.
Surface Pressure Distribution - Convert the measured pressure differences and free stream dynamic pressures to local (pressure coefficients). The log file format is defined in the procedure above.
From the result, select a value of δ and
is defined as the flow velocity at δ
Calculate the non-dimensional quantities and , where y is the vertical distance from the floor of the wind tunnel, and u is the flow velocity at y
Plot figures showing the chord-wise pressure coefficient distribution () on the airfoil vs. non-dimensional chordwise position, x/c, where c is the airfoil chord length. The standard in aerodynamics format is to plot as the ordinate with negative upward, and x/c as the abscissa.
Plot a figure showing the shape of the non-dimensional boundary layer profile. Plot as the ordinate vs. as the abscissa. Again, plot data points only (not connected with curves) and make sure you plot only one averaged point per traverse position.
There are two principal stresses for a cylindrical thin-wall pressure vessel (Figure 1): a longitudinal stress in the axial direction; and a hoop stressin the circumferential direction. This is a biaxial-stress state since there are two non-zero stress components. For a sufficiently long cylinder, we can relate these stresses to the difference between the internal pressure and the external pressure , sometimes referred to as the gauge pressure.
So ideally, the longitudinal stress if one-half the hoop stress for a cylindrical vessel, or .
For our geometry, we have , and . Thus we can simplify (7) to
Including our stress results for the ideal cylindrical vessel () gives
In this lab, you will record the wall strains in your pressure vessel as a function of time while the vessel is depressurized. The strains in the wall will go from one steady-state condition (fully pressurized) to another (depressurized). One way to characterize a transient system like this, moving from an initial state to a final state, is the “rise” or “fall” time, i.e., the time required for a chosen state variable to go from 10% to 90% of the difference from its final value, i.e., , where is the time required for state variable y to meet the condition . An example of this is shown in Figure 4.
Look up some common aluminum alloys used to manufacture soda cans. Find values for the modulus of elasticity, Poisson's ratio, and 0.2% offset yield stress for these alloys (you might try ).
Note the precise designation of the strain gauges you will use so that you can look up relevant parameters associated with this gauge such as the gauge factorand the transverse sensitivity factor. Note for T-rosette or rectangular gauges there is a value ofassigned to each grid.
If you are using a rectangular rosette gauge, you should bond the gauge in such a way that none of the 3 grids are aligned with the longitudinal and hoop directions.
Trigger the Vishay 7000 to start recording data. Grab the top of the can with two fingers while you open the can with your other hand. You should avoid squeezing the can during this step!Stop recording data and save the strain data to the hard drive.
The whole point is to apply the strain transformation formulas in order to recover the principle strain directions (maximum and minimum) along with the principle angle (qp). Once you calculate these three quantities you should of course recover the hoop and axial directions of the can.
You should see the dynamic strain change associated with the hoop and axial strain components on the live computer display. Your curves should feature a sharp change in strain, which then levels off at a steady value.
Gauge factors and transverse sensitivity factors for each of the strain gauges used in your test.
Estimate the dynamic strain rates by dividing the respective change in strain during the 10% to 90% interval by your rise time estimate. Do this for both the hoop and longitudinal strain signals.
Using the material database values of E and , use the Hooke’s Law Equations (8) to determine the changes in the can’s and corresponding to the strain changes determined in Step 5.
Using the appropriate results for and estimate, from Equations (3) and (5), the gauge pressure within the soda can before it was opened.
Plot of maximum principal stress angle as function of time for the can-opening experiment.
The plot should exhibit a chaotic and unstable variation during the first several seconds followed by a level, steady state value of . The “average” value of during the steady region should coincide with the expected direction for a hoop stress.
Table of estimated rise times (in units of µs) and strain rates (in units of) for both hoop and longitudinal strain signals.
Table of estimated can gauge pressure values (in psi) using each of your results for and (note, you will be reporting at least two pressure estimates).
Upper Tap Label | x/c | Lower Tap Label | x/c |
LE | 0 | L1 | 0.025 |
U1 | 0.025 | L2 | 0.1 |
U2 | 0.05 | L3 | 0.2 |
U3 | 0.1 | L4 | 0.3 |
U4 | 0.15 | L5 | 0.4 |
U5 | 0.2 | L6 | 0.5 |
U6 | 0.3 | L7 | 0.6 |
U7 | 0.4 | L8 | 0.7 |
U8 | 0.5 (DEAD) | L9 | 0.8 |
U9 | 0.6 | L10 | 0.898 |
U10 | 0.7 | TE | 1 |
U11 | 0.8 |
U12 | 0.9 |
One object of this experiment is to explore the displacement and strain behavior of structures using the digital image correlation technique. In addition, you will explore the interesting response of polypropylene, a material that exhibits different moduli in tension and compression. The experiment consists of two tests: 1) a four point bending test of a polypropylene specimen from which you will estimate the elastic moduli of polypropylene; and 2) a tension test of a second polypropylene specimen with a hole cut in it, from which you can determine the strain and stress concentrations caused by the hole. The loading for both tests will be accomplished using an Instron load frame.
In the context of structural testing, Digital Image Correlation (DIC) is a method for tracking the point-wise displacements of a structure (typically a surface of the structure) using a series of images of the structure undergoing deformation. DIC is a non-intrusive measurement technique since nothing has to be mounted to the specimen directly. Furthermore, DIC can measure real structural component geometries in real world conditions. The DIC measurements are primarily limited by image resolution, such that higher resolution images produce more accurate results. Alternatively, a higher-resolution displacement field can be captured by zooming the camera’s field of view to a smaller portion of the specimen of interest. To use DIC, it is usually necessary to prepare the specimen by painting a high-contrast speckle pattern on the surface so that subsequent pre- and post-deformation images can be analyzed to accurately determine the displacement field on the structural surface.
There are some requirements and rules-of-thumb for producing good speckle patterns. First, the pattern should be “random”; a highly organized and repeatable pattern would produce ambiguity in the displacement measurement. Second, the size of the speckles (or high contrast objects) should be roughly the same as (or slightly larger than) a 3x3 pixel region on the digital image for optimal tracking of the displacement. If the speckles are smaller than this, the image magnification can be changed to meet this criterion. Finally, the “density” of speckle features should be sufficient to have an average of 3-4 such features in a 10x10 pixel region. DIC does not rely on correlating a single speckle features, but rather multiple features to get an average displacement in a (multi-pixel) sub-region of the image.
DIC has its origin in speckle imaging approaches used in solid mechanics, and correlation-based analysis methods developed in the 1980’s for object tracking in image processing applications and particle-based velocimetry measurements in fluid mechanics. In fact, DIC is very similar to a common velocity-field measurement approach used in fluid mechanics, Particle Image Velocimetry (PIV). In both DIC and PIV, the individual displacements of many small subregions of an imaged area are obtained by comparing images before and after the displacement has occurred. For each subregion, the “before” (pre) and “after” (post) images are cross-correlated, sometimes using Fast Fourier Transform (FFT) algorithms. The displacement for that subregion is the one that provides the best correlation between the two images. In DIC, the displacement is the desired quantity. In PIV, this displacement is divided by the (short) time between the two images to obtain the local velocity.
This analysis process is typically performed after recording a sequence of images. In DIC, after the displacement field is calculated, the strain field can be determined. The two-dimensional surface displacement field is characterized as u(x,y) and v(x,y), where u and v are the displacements in the x and y directions for a point originally at location (x,y). With u and v determined, we can obtain the surface strain field using the strain-displacement relationships
(1)
Note that differentiating the displacement data amplifies the noise in the data; so advanced analysis software like the package used here employ additional processing approaches such as sophisticated smoothing to find the strain field from the displacement data.
In this lab, we will use the Aramis analysis software package.
You will also export the Aramis DIC data so you can analyze the full-field data.
Figure 1. The 3D DIC imaging systems, with two cameras mounted in a stereoscopic configurations and two light sources.
DIC (and PIV) systems come in different flavors. For example, the measured displacements of a thin region can be two-dimensional (2D) or three-dimensional (3D). In this lab, we will use a 3D DIC system (shown in Figure 1) that enables us to capture displacements in all three coordinate directions, including out-of-plane deformations. A 3D system provides more information than the more common 2D systems by adding an additional measurement. The 2D system requires only one camera (or equivalently, only camera imaging view point). To capture out-of-plane deflections, the 3D DIC system uses two cameras in a stereoscopic configuration. It is important to reiterate that 3D DIC (and PIV) systems provide three components of displacement (or velocity) from a surface (or thin planar region). There are also volumetric DIC and PIV approaches that provide 3D results for each location within a three-dimensional volume.
Our 3D DIC system, which employs two 5-megapixel cameras, will measure the displacements within a test volume that is centered on the test specimen. While the 3D DIC system provides accurate 3D displacement fields, it requires additional calibration effort compared to a simple, single-camera 2D system. The cameras are calibrated by orienting a thermally balanced plate with calibrated markings on it within the test volume.
The Aramis software will guide you through the calibration.
The four-point flexure or bending test is designed to test the flexural response of a slender beam. The goal of the four point bending test is to create a state of pure bending. Pure bending is a stress state where the bending moment is constant and the shear resultant is zero everywhere. The diagram on the left in Figure 2 illustrates a pure bending condition in which only opposing bending moments are applied to either end of the beam. Unfortunately, it is difficult to generate a pure bending moment in a real experiment. Instead, we will use opposing off-set point loads that generate a couple at either end of a beam. This configuration is shown on the right diagram in Figure 2. The advantage of this loading condition is that over the central span there are no shear loads and the beam is subject only to a bending moment.
Figure 2. Idealized pure bending load (left); four point bending test load (right).
Figure 3 illustrates the setup of the test apparatus for the four point bending test. The test specimen is placed on rollers which are placed below and above the specimen. The load frame applies a compressive load to the experimental apparatus which transmits point loads to the top and bottom of the beam. Rollers are used to ensure that simple support conditions are imposed.
Figure 3. Apparatus schematic for a four point bending test
Under ideal conditions in pure bending, the strain in the beam should be linear through the thickness:
(2)
Furthermore, if the material is linear elastic and isotropic, then Hooke's law applies and the bending moment can be calculated as follows:
(3)
where I is the second moment of area of the beam, which is given as
(4)
In Eq. (4), w is the width of the beam, and h is its depth. Therefore, if we impose M through a four point bending test, and can estimate from the digital image correlation results, we can infer the elastic modulus from
(5)
Polypropylene, however, exhibits different elastic moduli under tension and compression. As a result, the neutral surface is not at the geometric centroid of the cross-section and we have to use a composite beam analysis technique to find the elastic modulus under tension and compression. The measured strain will be offset from the geometric centroid, i.e., Eq. (2) has to be modified as follows:
(6)
where b is a strain offset.
The through-thickness stress distribution is shown in Figure 5. The y-location of the neutral surface can be found as . The bending moment can be found by a similar integration as that in Eq. (3), except using Eq. (6) for the strain and integrating separately on either side of the neutral surface, i.e.,
(7)
Integrating Eq. (7) yields
(8)
Substituting our expression for in (8) and simplifying gives
(9)
Next, we know from equilibrium that there cannot be an internal axial load in the beam. Therefore, the axial resultant (N) must be zero:
(10)
Integrating Eq. (10) and replacing with yields
(11)
Now combining Eq. (11) and Eq. (9) for the bending moment, eliminating __ and using our expression for the 2nd moment of area, Eq. (4), results in
(12)
Eq. (12) can be solved for the modulus under tension, giving
(13)
Inserting this into Eq. (11) provides an expression for the compressive modulus
(14)
Stresses around defects and sudden changes in a structure can be significantly higher than the average stress in the structure. These sharp increases in stress are called stress concentrations. A good example of a stress concentration is the behavior of the stress near a circular hole in a structure subject to uniform tension or compression. For an infinite plate loaded in-plane, the tangential stress around the edge of the hole has an analytic solution
(15)
where the coordinate θ runs circumferentially around the circular hole and is the average stress in the specimen far away from the hole. The direction θ=0 is aligned with the loading direction, or the x-axis in our geometry. The maximum value of the stress (due to the hole) normalized by is known as the stress concentration factor.
The stresses in planar cylindrical coordinates (i.e., , , and ) can be calculated from Cartesian stresses (i.e., , and ) using standard coordinate transformations, for example
(16)
(17)
In this lab, you will measure displacements and strains in a polypropylene test specimen (Figure 6) with a circular cutout subject to tension. The DIC system will enable us to visualize the distribution of strains around the hole and observe the stress and strain concentration.
Figure 4. Photograph of test specimen with hole.
A critically important step for a DIC experiment is to prepare the specimen's surface and to apply a speckle pattern that the software can interpret and transform into a strain field. Various resources such as Trilion's FAQ and Correlated Solution's technical article detail how best to do this. For this lab, the specimens have already been prepared, so the steps below are for your enlightenment.
If your specimen has a flat surface and good adhesion, self-adhesive labels with pre-printed speckle patterns can be directly applied to the specimen. The following process was followed to generate speckles for this lab:
Download the Correlated Solutions Speckle Generator software.
Convert the PDF to a high-resolution image using Adobe Acrobat Online or similar. This file can also be found below this list.
Open Dymo Connect, set up the Extra Large shipping label in portrait, and import the image from the previous step, scaling it up to full size.
Print labels.
Remove any previous labels and/or any large residues or deposits on the specimen surface.
Degrease the specimen's surface.
Carefully apply the label in the region of interest without folding the edges, and applying firm pressure to aid its adhesion.
Use a sharp knife to trim away the excess label, taking care to not cut away any from the region of interest.
For certain surfaces, such as those that aren't flat, speckling paint directly onto the specimen is the preferred approach. One such paint application method is using a toothbrush as follows:
Lightly sand and deburr the specimen to remove any manufacturing artifacts. Wipe clean the specimen to remove any oils/residues that may prevent paint adhesion.
Apply a matte white paint base layer to the specimen to remove any reflectivity it inherently has. Do this using spray paint, shaking the can well and applying in multiple thin layers with ample time for each layer to dry before applying the next.
Dip a toothbrush into black paint, tapping off any excess. Having determined the appropriate distance away from the specimen to get the desired particle size beforehand, flick the brush's bristles to create a distribution of speckles over the base layer that will give good results. Consult the DIC manual for recommended particle size and coverage.
Take a look at the tripod and ensure the legs are resting on the 3 X's marked on the floor
Visually inspect the camera and ensure it is facing the Instron machine with no yaw or pitch offset. This is VERY important since this is a 2D DIC system which cannot characterize out-of-plane displacements!
Ensure that the LED light power cord is plugged into the power strip.
Plug the USB cable from the Camera into the USB 3.0 port on the front of the computer. A small green light will illuminate on the back of the camera once you plug this cable in.
!! IF ANYTHING IS AWRY, DO NOT FIX IT YOURSELF. ASK YOUR TA FOR HELP !!
TA Note: Ask the Lab Manager or Head TA for assistance if necessary.
!! BE VERY CAREFUL NOT TO BUMP OR KNOCK THE CAMERA OR TRIPOD AROUND !!
It is best to load all specimens into the Instron machine from BEHIND the Instron machine.
Turn on the Instron Machine and the Instron PC if it is not already turned on.
Open Bluehill Universal and ensure the Software connects to the Machine.
Table 1. Four-point loading parameters to be used in experiment (defined in Figure 2)
Prepare the specimen:
Rotate the Instron grips such that it is facing forward if necessary. You will need 2 people to do this and make sure a TA is supervising this process!
Locate the square beam specimen that has a self-adhesive speckle pattern label applied.
Measure the specimen's thickness, width, and length.
Place the 4 point bending fixture into the Instron grips. You will need 2 people to do this and make sure a TA is supervising this process!
Grip the bending fixture such that the grip section on the fixture is flush to the back of the clamps. If done correctly, there should be a gap in the front.
It should also be gripped ~3mm below where the grip starts to diverge. See the image below for reference.
Locate the beam specimen that has a self-adhesive speckle pattern label applied.
Measure the thickness and width of the beam specimen being careful not to touch the speckle pattern with your fingers
With the help of a TA, ensure that there is no folding, fading, or any other damage to the speckle pattern. If there is, apply a new speckle pattern sticker to the specimen.
With the help of a TA, position the beam in the test fixture using a 12-inch lower support length and a 4-inch upper support length.
Place the rollers in locations to produce the four-point bending parameters listed in Table 1. There are arrows on the fixture itself to depict where the rollers must be placed
Ensure the bottom rollers are on placed below the beam. These 2 rollers are labeled with a "B"
Ensure the top rollers are placed above the beam. These 2 rollers are labeled with a "T"
TA Note: If this line starts to become faded, redraw it with the sharpie in the green toolbox at the end of the lab.
Jog the Instron Crosshead down until there is very little play in the top rollers.
You should use the fine adjustment wheel once you get very close to ensure you are not accidentally loading the specimen.
BE CAREFUL NOT TO PINCH YOUR FINGERS DURING THIS PROCESS!!!!
A good way to do this is to zero the load and stop and scroll the fine adjustment wheel back up AS SOON as you see a load being applied.
Another option is to use your fingers but again, be careful not to pinch your fingers.
Zero the load and displacement on the Instron software.
Verify LED Illumination:
Switch on the left side LED light by flipping the switch. Check to make sure it is illuminating the area of interest.
Switch off the left side LED.
Switch on the right side LED light by flipping the switch. Check to make sure it is illuminating the area of interest.
If both LEDs are pointing in the right direction, turn on both lights and keep them on.
!! IF ANYTHING IS AWRY, DO NOT FIX IT YOURSELF. ASK YOUR TA FOR HELP !!
TA Note: Ask the Lab Manager or Head TA for assistance if necessary.
Perform the experiment:
Open Vic-Snap 10 from the Start Menu
Under Project Path, select a save location that is convenient and easy to access.
Under Speckle prefix, type in "4pt" and under Calibration prefix, type in "4pt-cal"
Click OK. The software will open to the main screen.
You should now see the specimen on your screen.
If needed, HAVE A TA raise or lower the tripod with the hand crank.
On the bottom of the screen, you should see a slide rule where you can change the exposure time of the photograph. Slide this back and forth until you see no overexposure on the picture.
You can see overexposure when you see red dots on the image.
Your goal is to get as much exposure as possible without ANY overexposure ANYWHERE on the image.
When you are satisfied, click the "Focus" button on the top of the toolbar.
Verify that the predicted noise is below 0.05.
The camera has already been focused for you. If any changes are necessary, have your TA call the Lab Manager or Head TA for assistance.
Unclick the "Focus" button.
You are now ready to begin taking images.
Under Project on the left side of the screen, select the "Calibration" radio button.
This is the radio button directly underneath "Speckle"
Be sure not to select the Calibration tab!
Click the "Capture" button from the top toolbar.
Under Project and Under System on the left side of the screen, switch back to the "Images" tab. You should now see a number 1 next to Calibration indicating you have taken you calibration photograph
Select the "Speckle" radio button.
Take a reference image of the specimen WITHOUT any loading by clicking "Capture" from the top toolbar. Your count of Speckle photographs should now read 1.
Jog the Instron down using the fine adjustment wheel until you've applied a bending load of 150 lbf.
Click "Capture" from the top toolbar. Your count of Speckle photographs should now read 2.
Unload the specimen
Upload Images to Vic-2D:
Without, closing the Vic-Snap Software, open the Vic-2D 7 software from the start menu.
From the Home page, select the "Calibration images" button.
Locate the folder in which you have saved your images, and select the "4pt-cal" TIF file you captured earlier
Select "Speckle Images" button from the top toolbar (4th from the left)
Select and upload the Reference image and Loaded image TIF files you took earlier using Vic-Snap
On the left hand side under Project, you should now have 2 speckle images and one calibration image
Ensure that the proper image is selected as your reference image. It will have a red arrow next to the image.
If you are unsure which of your speckle images was the reference image, you can double click the image and see which one is under loading.
Calibrate the Scale
In this step, we will be using the calibration image to understand the dimensions of the specimen. We will draw a line across the dimension of the specimen that we know. Using this information, the software will be able to calculate strain in the dimensions we are interested in rather than pixels.
Click the "Calibrate Scale" button from the top toolbar (6th from the left)
Scroll in as best you can to zoom into the picture and pan to the required area. You can also resize the entire window if needed.
Once you are satisfied with how far you are zoomed in, select the "Manually Select" button under Tools (2nd from the left)
Holding the left click button on the mouse, drag a line across the entire width of the specimen
HINT: this line needs to be perpendicular to the length of the specimen. Think of the best way you can make this line as close to perpendicular as you can!
You can always redraw this line until you are satisfied
Under Point distance, enter the width of the specimen (1 inch). Click OK to save this. If you get prompted that this is a short line, select "Yes"
Perform Displacement Analysis
You should see a box that says "Aoi tools" near the top left of your screen. If you don't see this, select "Aoi Editor" from the "Window" drop down menu at the top of the screen
Select the "Create Rectangle" button under Aoi tools (2nd from the left)
Highlight your area of interest. This should be an area between the two rollers where you expect to see pure bending. You should also select a VERY small portion (a sliver) of the black area outside of the speckle pattern to ensure you are capturing the specimen edge effects!
You can delete and redo this as many times as you would like until you are satisfied with the area you select
Under Aoi tools, select the "Create Starting Point" button (2nd from the right)
Select a point somewhere within the area of interest.
Keep the default Subset value of 29 and Steps value of 7
Select the Green "Start Analysis" button from the top toolbar. Verify that the proper image is selected for your reference and that the loaded image is checked in the box below.
Also verify your output directory. Keep this the same as folder in which you saved your images so everything is in one place!
Click "Run"
You will see a preview of the analysis on the proceeding screen. Does it look like what you would expect to see? Click "Close"
Now that you have run your displacement analysis, we can calculate strain using this displacement data and the calibration image we characterized earlier.
Select the "Calculate Strain" button from the top toolbar (7th from the right)
Ensure that both ".out" files are selected for your computation. Click "Preview"
Does this look like what you would expect?
Discuss!
If you are satisfied with this preview and are confident that this looks like what you would expect, click "Start"
Click "Close"
Visualize your Results
Under the "Plot" drop down menu at the top of the screen, select "New Plot"
Under "Plotting Tools" on the left hand side, make sure you are on the "Contour" tab
Under "Project" on the left hand side, go to the "Data" tab. Double click your loaded image.
Under "Plotting Tools" on the left hand side again, change the variables between xx, yy, xy, and principal strains and observe all the plots. You can and should also observe displacements in all 3 cartesian directions
Discuss what you see with your labmates!
Take notes in your lab notebooks!
Does this all make sense?
Discuss with your TA
Redo any portions of the procedure if needed
If you are satisfied with your results, select the "Data" drop-down menu and under "Export" select "All Data"
Make sure both of your ".out" files are selected and click "Export"
You may select any file format that works for you. The default is a CSV and is what we recommend.
Prepare the specimen:
Rotate the Instron grips such that it is facing sideways if necessary. You will need 2 people to do this and make sure a TA is supervising this process!
Locate the open-hole dog-bone specimen that has a self-adhesive speckle pattern label applied.
Measure the specimen's thickness, width, and hole diameter.
With the help of a TA, ensure that there is no folding, fading, or any other damage to the speckle pattern. If there is, apply a new speckle pattern sticker to the specimen.
With the help of a TA, position the specimen in the clamping test fixture.
Zero the load and displacement on the Instron software.
Verify LED Illumination:
Switch on the left side LED light by flipping the switch. Check to make sure it is illuminating the area of interest.
Switch off the left side LED.
Switch on the right side LED light by flipping the switch. Check to make sure it is illuminating the area of interest.
If both LEDs are pointing in the right direction, turn on both lights and keep them on.
!! IF ANYTHING IS AWRY, DO NOT FIX IT YOURSELF. ASK YOUR TA FOR HELP !!
TA Note: Ask the Lab Manager or Head TA for assistance if necessary.
Perform the experiment:
Open Vic-Snap 10 from the Start Menu
Under Project Path, select a save location that is convenient and easy to access.
Under Speckle prefix, type in "OpenHole" and under Calibration prefix, type in "OpenHole-cal"
Click OK. The software will open to the main screen.
You should now see the specimen on your screen.
If needed, HAVE A TA raise or lower the tripod with the hand crank.
On the bottom of the screen, you should see a slide rule where you can change the exposure time of the photograph. Slide this back and forth until you see no overexposure on the picture.
You can see overexposure when you see red dots on the image.
Your goal is to get as much exposure as possible without ANY overexposure ANYWHERE on the image.
When you are satisfied, click the "Focus" button on the top of the toolbar.
Verify that the predicted noise is below 0.05.
The camera has already been focused for you. If any changes are necessary, have your TA call the Lab Manager or Head TA for assistance.
Unclick the "Focus" button.
You are now ready to begin taking images.
Under Project on the left side of the screen, select the "Calibration" radio button
This is the radio button directly underneath "Speckle"
Be sure not to select the Calibration tab!
Click the "Capture" button from the top toolbar.
Under Project and Under System on the left side of the screen, switch back to the "Images" tab. You should now see a number 1 next to Calibration indicating you have taken you calibration photograph
Select the "Speckle" radio button.
Take a reference image of the specimen WITHOUT any loading by clicking "Capture" from the top toolbar. Your count of Speckle photographs should now read 1.
Jog the Instron crosshead up using the fine adjustment wheel until you've applied a tensile load of 2.2 kN.
Click "Capture" from the top toolbar. Your count of Speckle photographs should now read 2.
Unload the specimen
Upload Images to Vic-2D:
Without, closing the Vic-Snap Software, open the Vic-2D 7 software from the start menu.
From the Home page, select the "Calibration images" button.
Locate the folder in which you have saved your images, and select the "OpenHole-cal" TIF file you captured earlier
Select "Speckle Images" button from the top toolbar (4th from the left)
Select and upload the Reference image and Loaded image TIF files you took earlier using Vic-Snap
On the left hand side under Project, you should now have 2 speckle images and one calibration image
Ensure that the proper image is selected as your reference image. It will have a red arrow next to the image.
If you are unsure which of your speckle images was the reference image, you can double click the image and see which one is under loading.
Calibrate the Scale
In this step, we will be using the calibration image to understand the dimensions of the specimen. We will draw a line across the dimension of the specimen that we know. Using this information, the software will be able to calculate strain in the dimensions we are interested in rather than pixels.
Click the "Calibrate Scale" button from the top toolbar (6th from the left)
Scroll in as best you can to zoom into the picture and pan to the required area. You can also resize the entire window if needed.
Once you are satisfied with how far you are zoomed in, select the "Manually Select" button under Tools (2nd from the left)
Holding the left click button on the mouse, drag a line across the entire width of the specimen
HINT: this line needs to be perpendicular to the length of the specimen. Think of the best way you can make this line as close to perpendicular as you can!
You can always redraw this line until you are satisfied
Under Point distance, enter the width of the specimen (1.5 inches). Click OK to save this. If you get prompted that this is a short line, select "Yes"
Perform Displacement Analysis
You should see a box that says "Aoi tools" near the top left of your screen. If you don't see this, select "Aoi Editor" from the "Window" drop down menu at the top of the screen
Select the "Create Rectangle" button under Aoi tools (2nd from the left)
Highlight your area of interest. You should select the area around the open hole all the way to the edge of the specimen.
Select a vertical length that is about 1.5x the width of the dogbone specimen, creating a rectangle
You should also select a VERY small portion (a sliver) of the black area outside of the speckle pattern to ensure you are capturing the specimen edge effects!
You can delete and redo this as many times as you would like until you are satisfied with the area you select
Select the "Cut Circle" button under Aoi tools (3rd from the right)
Cut out dark section of the hole MAKING SURE you are not cutting out any part of the specimen itself!
If necessary, select a starting point by clicking "Select Starting Point" under Aoi tools (2nd from the right).
Keep the default Subset value of 29 and Steps value of 7
Select the Green "Start Analysis" button from the top toolbar. Verify that the proper image is selected for your reference and that the loaded image is checked in the box below.
Also verify your output directory. Keep this the same as folder in which you saved your images so everything is in one place!
Click "Run"
You will see a preview of the analysis on the proceeding screen. Does it look like what you would expect to see? Click "Close"
Now that you have run your displacement analysis, we can calculate strain using this displacement data and the calibration image we characterized earlier.
Select the "Calculate Strain" button from the top toolbar (7th from the right)
Ensure that both ".out" files are selected for your computation. Click "Preview"
Does this look like what you would expect?
Discuss!
If you are satisfied with this preview and are confident that this looks like what you would expect, click "Start"
Click "Close"
Visualize your Results
Under the "Plot" drop down menu at the top of the screen, select "New Plot"
Under "Plotting Tools" on the left hand side, make sure you are on the "Contour" tab
Under "Project" on the left hand side, click the "Data" tab. Double click your loaded image.
Under "Plotting Tools" on the left hand side again, change the variables between xx, yy, xy, and principal strains and observe all the plots. You can and should also observe displacements in all 3 cartesian directions
Discuss what you see with your labmates!
Take notes in your lab notebooks!
Does this all make sense?
Discuss with your TA
Redo any portions of the procedure if needed
If you are satisfied with your results, select the "Data" drop-down menu and under "Export" select "All Data"
Make sure both of your ".out" files are selected and click "Export"
You may select any file format that works for you. The default is a CSV and is what we recommend.
CHECK YOUR DATA FILES. REDO ANY PORTIONS OF THE LAB YOU FEEL YOU MUST REDO
Close the Vic-2D and Vic-Snap softwares. You may save the project to your folder in Vic-2D if you wish but it is not necessary
Turn off both the LED lights by flipping the switches.
Remove the specimens from the Instron Machine and place them on the Table for the next groups.
Unplug the Camera USB cable from the front of the computer.
Have your TA upload your data to Canvas. It is best to zip your entire folder containing all your images, output files, and exported data files
Thickness and width of the beam, and load value used in the four-point bending test.
Displacement and strain data from the VIC-2D software for the four-point bending test.
Thickness and width of the specimen, hole diameter, and load value used in the open-hole tension test.
Displacement and strain data from the VIC-2D software for the open-hole tension test.
Using a best linear fit for the appropriate region of your plots from the four-point bending test, compute the slope () and intercept (b) relative to a coordinate axis centered on your specimen.
From your experimental data, determine the tensile and compressive moduli ( and ) for polypropylene.
From your plots of transverse normal strain divided by the axial strain , determine a value for Poisson’s ratio () for polypropylene.
For the open-hole tension test, compute the stresses in your specimen based on the measured strains and the appropriate measured modulus of elasticity for polypropylene.
Table of beam dimensions and load value used in the four-point bending test.
Table of specimen dimensions, hole size and load value used in the open-hole tension test.
“Image graphs” or “full-field color plots” of the axial (horizontal) and transverse (vertical) displacement field for the field of view analyzed by the VIC-2D software in the bending test. Choose your color scaling wisely to accentuate any important gradients. Be sure to also include color bars showing how your colors map to the values of displacement.
Here an “image” means a false-color image, where each displacement (or strain) measurement location is like a pixel in the false-color image, and the color of the pixel corresponds to the value being shown (axial displacement in this case). You can do this, for example, with the image function in Matlab
Images of , and for the field of view imaged and analyzed by the VIC-2D software in the bending test, with color bars.
For the four-point bending test, plots of the axial and transverse normal strain fields as a function of y (with the y-axis as defined in Figure 5), at three axial locations between the two inner load points (i.e., within the constant moment region). One of these should be the center axial location.
Figure 5. Through-thickness stress-strain distribution for material with different moduli in compression and tension.
For the bending test, plot of the transverse normal strain divided by the axial strain along the center line used for Result 5.
Table of measured values of , b, and the polypropylene properties , , and . Include in your table the published values for the polypropylene properties (include your source for those values).
Images of the axial strain () and transverse strain () fields for the region analyzed by the VIC-2D software in the open-hole tension test, with color bars.
Images of the axial () and transverse stress () fields based on your measured strains and measured modulus of polypropylene, with color bars for the open hole tension test.
Images of the normal radial stress () and normal tangential stress () fields for the open-hole tension tests, with color bars.
A single graph containing plots of the tangential stress along a line perpendicular to the length dimension of your specimen that passes through the center of the hole for the open hole tension test.
A single graph containing a plot of the tangential stress (normalized by its value far from the hole) as a function of angle around the hole, close to the edge of hole.
This laboratory introduces the measurement of gas temperature and explores the performance of a jet engine. First, the operation and proper use of thermocouples is explored. Then, thermocouples are used to measure gas temperatures at various locations in the jet engine. In addition, engine pressures are measured with piezoresistive transducers.
NOTE: All combustion experiments are potentially hazardous. Please follow all precautions outlined in the safety section and given during the lab.
There are a number of devices used for the measurement of temperature. Most rely on the effect temperature has on some other measurable property of a substance. One of the oldest types is the expansion thermometer, which relies on the change in volume of a substance (usually a liquid) as the temperature varies. An example is the glass tube, mercury-filled thermometer, in which the height of the mercury column becomes a measure of temperature. While easy to manufacture, expansion thermometers are not easily implemented as small, compact electronic sensors.
Various electronic temperature sensors have been developed, including thermocouples, resistance thermometers and quartz resonance thermometers. In a quartz resonance thermometer, temperature is determined by its influence on the resonant frequency of a quartz crystal. Since frequency can be determined to great accuracy, these devices have potential for high resolution. Resistance temperature detectors (RTDs) rely on the change in resistance of a material with temperature; thus measurement of resistance is converted to a temperature measurement. There are two basic types of RTDs: metal based devices and semiconductor devices (also known as thermistors). While quite accurate temperature measurements can be obtained with RTDs and crystal thermometers,* they are usually limited to “low temperature” operation. For example, platinum resistance thermometers can operate only up to ~1000°C, while thermistors are generally limited to a few hundred degrees Celsius. Thermocouples, on the other hand, can provide reasonably accurate measurements at temperatures up to at least 2600 K with proper choice of materials.
*Quartz resonance thermometers and RTDs can have precisions better than 10-4 °C.
Figure 1. Simple (open) thermoelectric thermocouple circuit.
Thermocouples rely on the voltage produced by a temperature difference between two junctions formed between thermoelectrically dissimilar metals (see Fig. 1). In other words, a thermocouple is simply two different types of metals, usually in the form of wires, connected together. The voltage is produced because a temperature gradient in a metal conductor also induces a gradient in electron density in line with the temperature gradient. It can be shown that the voltage produced between the two junctions of the Fig. 1 circuit is given by:
(1)
where T(x) is the temperature distribution along each wire, and ε is called the thermoelectric power of the material. Thus it can be seen that the voltage difference is generated throughout the length of the wires, and is due to the local temperature gradient.
Equal to the sum of the Thomson coefficient and temperature derivative of the Peltier coefficient for the metal.
When the wire is perfectly uniform in composition, such that ε is not a function of position, and the two wires are connected between and , the integrals in equation (1) become,
(2)
In this case, one can think of the voltage produced in a thermocouple as strictly due to the temperature difference between the junctions. It is important to remember that this holds only for the above uniformity assumption. In fact, some descriptions of thermocouples erroneously state that the voltage is produced “at the junction”, when in fact it is produced wherever there is a temperature gradient in the metal. The thermocouple circuit that will be considered the “ideal” circuit is shown in Fig. 2. The difference between it and the circuit of Fig. 1 is simply that the voltage is “measured” at the reference junction, instead of midway through one of the legs of the thermocouple. Note, both points in the reference junction must be at the same temperature (isothermal) to be equivalent to the circuit of Fig. 1.
Figure 2. Ideal thermoelectric circuit, with the thermocouple voltage measured across the isothermal reference junction.
While any two metals with different ε can be used to produce a thermocouple,** a small number of metals (both pure and alloys) have been identified for their stability, linearity, reproducibility, and high temperature capability. Table 1 lists some common pairs of thermocouple materials, including their approximate limiting operating temperature. Some of these are sufficiently common that they are considered standards and are denoted simply by a letter. For example, a chromel/alumel device is called a “type K” thermocouple, and it has a nearly linear temperature sensitivity (see Fig. 3). The pairs of metals listed in Table 1 were also chosen for their good temperature sensitivity, which is normally achieved by picking materials that have ε with different signs. Figure 4 shows the voltage that would be generated along a single, homogeneous wire of various materials as a function of the temperature difference between its two ends. For the type K thermocouple, the chromel and alumel alloys produce voltages of opposite sign for the same temperature gradient.
**In fact, there have been a number of applications where the operating temperature of a machinery part has been measured using the machine structure itself as part of the thermoelectric circuit.
Chromel is a nickel-chromium alloy; alumel is a nickel-aluminum alloy
*Limited use in vacuum or reducing environments Limited use in oxidizing at high temperature
Table 1. Some standard thermocouple materials and their properties.
Figure 3. Sensitivity of a standard type-K thermocouple, based on ITS-90 inverse polynomial fit. The millivolt output is based on a reference temperature of 0°C.
The temperature behavior shown in Fig. 4 can be used to explain what happens in a thermocouple circuit like that of Fig. 2. Consider a type-J thermocouple connected between a reference junction at and a junction at higher temperature, . As shown in Fig. 5, the voltage that would be produced is found by starting at the reference temperature, following a line with a slope equal to up to , and then switching to a line with the slope of back to . Thus the constantan end of the reference junction will be at a higher voltage than the iron. If the temperature at the probe junction is raised, the thermocouple voltage increases. The reason that two dissimilar materials must be used is also evident in Fig. 5. If iron was used for both legs of the thermocouple, the voltage developed in the first leg of the circuit would be canceled as the temperature drops in the other leg, i.e., we would follow the iron curve upward to , and then back down to with no net voltage induced.
Figure 4. Voltage−temperature response for several metals.
Figure 5. Development of the voltage difference across the thermocouple, Vth, in a thermocouple circuit, like that of Fig. 2, formed by two metals.
While the circuit of Fig. 2 is ideal, it is usually not practical. First, thermocouple wire can be somewhat expensive. Thus it becomes costly if the measuring device needs to be located remotely from the experiment. Also, it is not uncommon to connect multiple thermocouples to a single measurement device through some sort of switching circuit. Finally, the measurement device may have wires and connectors of its own, which effectively become part of the thermocouple circuit. Therefore, it is important to consider modifications to the ideal circuit. For example, Fig. 6 shows a circuit in which the reference junction is connected to the measurement device through copper wires. As seen in Fig. 7, the voltage difference across the device junction () is identical to the reference junction voltage () if the copper wires are identical, and if the device connections (1 and 5) are both at the same temperature. In this case the voltage developed in leg 1-2 is counteracted by the voltage produced in leg 4-5.
Figure 6. Modified circuit; copper extension wires connect the thermocouple to the measuring device, and the device connections are at the same temperature.
Figure 7. Development of the thermocouple voltage difference for the thermocouple circuit shown in Fig. 6. If the wires connecting points 1 to 2, and 4 to 5 are identical, and the temperatures and are the same, then the voltage measured across the device () is equal to the voltage that would be produced by the ideal thermocouple circuit ().
So far we have explored the thermocouple voltage that is produced between a junction at an unknown temperature and another junction at . To convert the thermocouple voltage to the unknown requires us to know two things: 1) the change in voltage associated a given temperature change, i.e., the temperature sensitivity of the thermocouple materials, and 2) .
First, consider ways to determine . There are two basic approaches: 1) create a situation where is fixed by some physical condition, and 2) measure with another device. A known temperature can be produced using a phase point of a material, for example 0°C can easily be produced to within 0.01°C accuracy by a bath of liquid water and ice, if the ice and water are both present and allowed to come to equilibrium (this generally requires crushing the ice to small size and putting the mixture in an insulated container). If the reference junction (or points 2 and 4 in Fig. 6) are place in the ice bath, then will essentially be 0°C. It many situations, however, it is impractical to require access to ice. Therefore, a popular approach is to measure the reference junction’s temperature with a thermistor or similar device that can provide accurate, absolute temperature measurements, though at low temperatures.
With known, all that remains is to convert the thermocouple voltage to temperature. Standard thermocouple materials have been extensively studied at the National Institute of Standards and Technology (NIST), and the voltage produced by a thermocouple at Thot is generally reported for , and the values are available in tables, graphs (such as Fig. 3) or polynomial fits (see Table 2). If one is using a 0°C reference junction, then you simply look up the measured thermocouple voltage in the table (or graph or fit) and find the corresponding temperature. If you are using a different reference temperature, you must first add an offset voltage to your measured voltage. The offset is the voltage would be produced by a thermocouple at your measured referenced to 0°C. Electronic ice reference circuits exist to do just this, the add the proper offset voltage to account for .
Table 2. ITS-90 thermocouple inverse polynomials for type K thermocouples; two polynomial fits are listed, for separate temperature/voltage ranges. The reference junction is assumed to be at 0°C, and the polynomials are of the form, , with T in degrees Celsius and V in microvolts.
Strictly speaking, a thermocouple sensor measures the temperature of the thermocouple junction itself, which of course is not what we usually want to know. Rather, we wish to determine the temperature of the body in which the thermocouple is embedded. For gas temperature measurements, the “body” of interest is the gas. As shown in Fig. 8, there are two basic thermocouple probe arrangements: one in which the thermocouple junction is immersed in the gas (unshielded probe); and one where the junction is inside some housing material (usually a metal), and the housing is immersed in the gas (shielded probe). The former provides a better measure of the gas temperature and a better time response; the latter approach protects the thermocouple from damage or exposure to incompatible gases (see Table 1). For high speed flows, there are also shielded stagnation probes, which are designed to slow the flow down to a very low velocity before it contacts the thermocouple junction (see Fig. 8 for a simple example).
Figure 8. Examples of unshielded, shielded and stagnation thermocouple probes.
In general, the thermocouple temperature can not exactly equal the gas temperature due to heat losses. Assuming the gas is the hotter material, it will heat up the initially colder thermocouple junction. If the thermocouple had no way of losing energy, then it would eventually heat up to the gas temperature. However, the thermocouple can lose heat; either by thermal conduction from the junction down through the thermocouple wires, or by radiation. Therefore even in steady-state operation, the thermocouple will tend to be at a lower temperature than the surrounding (hot) gas in low speed flows. For high speed flows, the thermocouple temperature will also be affected by the conversion of the flow’s kinetic energy to thermal energy in the region in front of the probe. Therefore in high speed flows, the thermocouple temperature will generally exceed the freestream static temperature.
You will also be making pressure measurements in this lab with a transducer that is something like the Barocel/Baratron type transducers used in a previous lab, i.e., a differential pressure is determined from the movement of a thin diaphragm exposed to the pressure difference. Instead of the capacitance based approach of the Barocel/Baratron devices, the sensors used in this lab consist of a miniature diaphragm and strain sensors composed of semiconductor material and manufactured using MEMS (Micromachined Electro-Mechanical Systems) technology. The sensors are described in more detail in a following lab.
In most cases, with the exception of civil aviation, modern aircraft are powered by turbine engines (also called gas turbines). While piston engines are efficient for low power applications, their power (or thrust) to weight ratio drops significantly as engine power increases. This makes them unsuitable for large aircraft that require high power engines. Gas turbine engines are also used in a number of other applications including marine propulsion, operating gas pipeline compressors, and most notably, the generation of electric power.
Figure 9. Basic components of a gas turbine engine.
Figure 10. Basic components of a gas turbine engine.
The components of a basic gas turbine system are shown in Fig. 9. Air enters the engine (through an inlet not shown) and passes through a rotating compressor that raises the air pressure. Next, the high pressure air enters the combustor where it is mixed with fuel and burned without much change in pressure. The hot products then pass through a rotating turbine that extracts work from the flow and sends it to the compressor via a rotating shaft. The exhaust of the turbine is a hot, high pressure gas. In a turbojet, the exhaust is expanded through a nozzle (Fig. 10), which converts the thermal energy to kinetic energy, i.e., it accelerates the gas in order to produce thrust. On the other hand in a turboshaft engine, the hot, high pressure gas exiting the first turbine is expanded through a following power turbine that converts the thermal energy to shaft power, which can be used to run a rotor in a helicopter engine or to turn an electric generator in an aircraft’s auxiliary power unit (APU) or in a ground power station. In this lab, you will operate and perform measurements on an SR-30 turbojet manufactured by Turbine Technologies, Ltd. (Fig. 11), operating on Jet-A fuel.
Figure 11. Schematic of the gas turbine system to be used in this lab.
The following is a brief description of the thermodynamic expressions used to analyze a jet engine.*** The processes that occur in a turbine engine can be modeled, in the ideal case, by a Brayton cycle. As shown in Fig. 12, the ideal compressor (2→3), turbine (4→5) and nozzle (5→e) can be modeled as isentropic processes (constant entropy, s). The ideal combustor (3→4) is modeled as a constant pressure heat addition, where the “heat” comes from burning the fuel, and the heat release rate is given by
(3)
where is the fuel mass flow rate and HV is the heating value of the fuel (~for most liquid jet fuels).
***A more detailed development can be found in the text for AE 4451, Hill and Peterson’s Mechanics and Thermodynamics of Propulsion.
Meaning each component of the engine is reversible and has no heat losses
Figure 12. Ideal Brayton cycle; stations 2-e correspond to the numbering scheme shown in Fig. 10.
For the three isentropic processes, we can find a relationship between the ratios of temperature (absolute, i.e., in Kelvin or Rankine units) and pressure (absolute, not gauge pressure) across each device from the entropy state equation for a perfect gas, i.e., for an isentropic process going from state a to state b,
(4)
Assuming that the gas is also calorically perfect (=constant),
(5a)
or equivalently for stagnation properties,
(5b)
where is the ratio of specific heats ().
There is also a relationship between the stagnation temperature change across the compressor and turbine. Since the turbine is used to power the compressor, the output power of the turbine should equal the power input to the compressor (assuming steady operation and no shaft losses, which are typically less than 1% of the shaft power). Therefore from conservation of energy (and for adiabatic conditions),
(6)
where is air mass flow rate entering the engine, and and are the (average) specific heats of the gases passing through the turbine and compressor. Similarly, the velocity at the nozzle exit can be found from
(7)
where is the (average) specific heat of the gas passing through the nozzle, and (again assuming the nozzle is adiabatic, i.e. no heat losses). For the combustor, again using energy conservation, we can calculate the expected change in temperature caused by burning the fuel:
(8)
where f is the fuel-air ratio .
If the compressor, turbine and nozzle are not ideal, i.e., they are not reversible (but still adiabatic), the temperature and pressure ratios across each component are related by the adiabatic component efficiencies:***
(9)
(10)
(11)
Note, if the nozzle is truly adiabatic, then .
We can also define an overall thermal efficiency of a static (stationary) turbojet engine, which is given by,
(12)
If the turbojet is ideal (and is assumed constant throughout the engine), it can be shown that the thermal efficiency should solely be a function of the cycle pressure ratio:
(13)
Thus as pressure ratio of the compressor increases, of the engine should increase.
As with any combustion experiment safety is a primary concern. Improper operation of the jet engine can damage the engine, and pose a hazard to those nearby. Carefully follow all the safety instructions presented during the lab concerning startup, operation and shutdown of the engine. Do not operate the engine without direct supervision by the lab technician. Do not place any part of your body, or any objects in the inlet or exhaust regions. All students working in the lab MUST WEAR safety glasses/goggles in the lab and hearing protection when the engine is operating (both will be supplied to you).
The following items must be turned in at the start of your lab session.
Using the information supplied in Table 2, determine the thermocouple voltage you would expect to measure if the type-K thermocouple junction was at room temperature (~74°F) and the reference junction was placed in an ice bath.
Bring a modified copy of the equation you developed in the earlier labs that relates dynamic pressure to velocity of a gas, and the ambient pressure and temperature. You need to modify the equation to use dynamic pressure in psi rather than mm Hg.
For this experiment you will need the items in the image below, where:
A = Digital weather station - used to measure atmospheric temperature and pressure
B = Anti-spill tumbler holder - once you've filled the thermal tumblers, place them in here to prevent them tipping over and spilling
C = Thermal tumbler for iced water - fill with ice and tap water from the combustion lab kitchen
D = Thermal tumbler for boiling water - fill with boiling water from the combustion lab kitchen kettle
E = Thermocouple-to-multimeter adapter (all copper wires)
F = Reference junction breakout cable (copper wires between multimeter and terminal block, K-type thermocouple extension wire (also known as KX wire) between terminal block and thermocouple - this is a physical realization of the schematic in Figure 6 but with alumel and chromel)
G = Cold junction compensator - an electronic device that acts as a surrogate ice bath reference
H = K-type shielded thermocouple
J = Digital multimeter
Note that with all of these connectors and junctions, there is a reasonable chance that throughout the week at least one will loosen and allow a flaky electrical connection. If this occurs, check through the wiring and debug where any loose connections are, tightening any loose screws/connectors. The thermocouple connectors may be opened to reveal the screw terminals inside.
Also note that you should only record multimeter readings when the thermocouple output has stabilized. This can take a little time due to the thermal mass of the shielded thermocouple.
Locate the storage box containing all the items from the list above and prepare the iced and boiling water in the combustion lab kitchen. Be sure to keep the lids sealed tight to minimize temperature change, and to place the tumblers in the anti-spill tumbler holder.
Using the electronic weather station, record the ambient pressure and temperature.
Connect the thermocouple and the multimeter to the reference junction breakout cable (black cable to multimeter -COM port, red cable to +V.mA.Ω port).
With the multimeter set to read DC voltage ("°C/°F V" setting), record the voltage with the thermocouple exposed to:
Ambient air
Iced water
Near-boiling water
With the thermocouple exposed to the ambient air, expose each of the terminal block junctions to the water baths. You should do this four times:
Where only the wires connected to the thermocouple are (reference junction)…
In the ice water
In the near-boiling water
Where only the wires connected to the multimeter are (device junction)…
In the ice water
In the near-boiling water
Remove the multimeter and thermocouple from the reference junction breakout cable.
Connect the thermocouple directly to the cold junction compensator, and the cold junction compensator to the multimeter via the thermocouple-to-multimeter adapter.
Having switched on the cold junction compensator, record the voltage with the thermocouple probe exposed to:
ambient air
iced water
near-boiling water
Close up the experiment:
Turn off the cold junction compensator and multimeter.
Disconnect all components.
Pour out the tumblers and return to the holder with the lids off to air dry.
Return everything to its storage box.
Proceed to the turbine engine experiment.
The turbine engine can be extremely dangerous if not operated correctly. Follow the TAs instructions at every step, wear your ear and eye PPE, and never stand in front of the inlet port when running
Take a tour of the turboshaft engine cutaway in the Combustion Lab foyer.
Return to the lab turbojet engine, the Turbine Technologies SR-30, and have the TAs help open the blast hood. You will need to un-brake the cabinet wheels, move it slightly so the extraction duct is clear, and then reapply the brakes.
Familiarize yourself with the flow path through the engine (the poster on the engine inlet and the LabView VI, as well as documentation from the manufacturer, may help also). Also observe how the throttle handle on the control panel varies engine throttle.
With the electrical power and compressed air switched off, and with no one stood next to the control panel to accidentally switch something on:
Rotate the compressor by hand to check that it freely rotates. Have anyone who wants to do this also have a turn.
Break your group into 2 sub-groups at random. Using the measuring tools provided, each person in the first sub-group will take turns measuring the outer diameter of the shaft, with each person in the other group measuring the inner diameter of the inlet. With the guidance of the TAs, compare individual results and select a final measurement for each dimension. These will be used to determine the cross-sectional area of the inlet.
Remove any foreign objects, replace the blast hood, and move back into place with the extractor slightly inside the hood and the wheel brakes applied.
Review the engine startup and shutdown procedure with the TAs (embedded below this list, and printed out on the engine control panel and by the PC). Familiarize yourself with the emergency exit routes, fire extinguisher locations, and emergency shutdown procedure.
Open the LabView VI and prepare it for recording:
Hit run and choose a save location.
Input the room temperature and pressure as read from the electronic weather station (beware of units!).
Sanity check that the live readings make sense (the TAs will help):
RPM will not be valid until running.
Temperatures should be similar to room temperature.
Pressures should be close to zero.
Fuel flow rate should be around 0.5V.
Hit "Record Data" to save a data point in non-running conditions.
Familiarize yourselves with Step 13 below for what will need to happen by way of data gathering when the engine is running.
Have everyone stand around the sides of the engine so that they are both protected by the polycarbonate shields of the enclosure, and can see either the large monitor or the control panel. No one should stand beyond upstream of the inlet!!!
Designate one person to press the "RECORD DATA" button at each data point. Have them stand near the engine operator to co-ordinate timing.
Do all the steps in the startup procedure (taped to the engine control panel, and shown here below) right up to starting the engine.
Having put on your PPE and got your video camera ready (for fun, not for a deliverable!), have the TA start the engine.
Once the engine is warmed up, acquire data as follows:
Capture data at roughly the following RPMs:
Idle
35,000
40,000
45,000
50,000
55,000
60,000
Full throttle
RPM lags throttle so allow the RPM to settle before taking a reading.
To minimize running time and thus fuel consumption, don't worry about hitting the above RPM targets perfectly... just get it within 1000 RPM or so and capture data.
DO NOT EXCEED 82,000 RPM.
When you are finished, have the TA shutdown the engine according to the "Initial shutdown sequence".
Click "Done" in the LabView VI. Check the data is saved successfully by having the TA import it into MS Excel and compare the approximate order of magnitude of results with known good data.
If the data is not satisfactory, re-start the engine and repeat Step 13 onwards. If it is satisfactory, email the data to yourself or upload it to Canvas, then help the TA follow the "Full closedown sequence" procedure to turn off the compressed air, exhaust, etc.
Ambient pressure and temperature.
Voltages from the type-K thermocouple probe connected to the voltmeter under the different connection and temperature cases described above.
Cross-sectional area of the compressor inlet.
Values of:
dynamic pressure and at the compressor inlet
and (gauge) at the compressor exit
(gauge) and at the turbine inlet
(gauge) and at the turbine exit
(gauge) and ~ at the nozzle exit
pressure drop (a voltage signal) in the fuel system at various RPM settings.
Thermocouple probe temperatures reduced from the measured voltages (without heating or cooling the strip terminal junctions).
Air mass flowrate (from dynamic pressure at the compressor inlet).
Fuel mass flowrate (from fuel pressure reading and calibration, see Table 3).
Table 3. Calibration data for fuel flowrate
Compressor efficiency.
Exit static temperature determined from the nozzle exit (which is measured as a gauge pressure here) and the exit stagnation temperature .
Nozzle exit velocity.
An estimate of the heat loss rate from gas to the nozzle (based on the difference in stagnation temperature across the nozzle).
Compressor and turbine powers (assuming adiabatic operation).
The actual engine thermal efficiency.
Tables of the thermocouple voltages and reduced temperatures for the different cases.
Tables of the raw and reduced engine conditions at each operating RPM.
Plot of the compressor pressure ratio versus air mass flowrate.
Plot of the compressor efficiency versus air mass flowrate.
Plot of the heat loss in the nozzle rate versus air mass flowrate.
Plot of the engine thermal efficiency versus compressor pressure ratio.
Philip Hill and Carl Peterson, Mechanics and Thermodynamics of Propulsion, 2nd edition, Prentice-Hall, 1992.
This lab covers the use of static taps for the measurement of static and stagnation pressures, area ratio, and Mach number in nominally steady, compressible (subsonic and supersonic) flows. In addition, it includes the use of the Schlieren technique to visualize shock and expansion waves in supersonic flows. The various measurements are acquired in an intermittent supersonic wind tunnel that utilizes a converging-diverging nozzle. The experiments also allow the student to experimentally investigate many interesting and important aspects of the behavior of supersonic flows.
The supersonic wind tunnel used in this experiment is the AF300 Intermittent Supersonic Wind Tunnel designed by TecQuipment. This is a two-dimensional wind tunnel, meaning that the test articles span the width of the entire test section. It has the capability of accelerating flow to three different speeds: subsonic, Mach 1.4, and Mach 1.8 depending on the liner that is put into the wind tunnel. A picture of the wind tunnel is shown in Figure 1 below.
The wind tunnel works using the concept of induction. A downstream flow of high pressure compressed air induces an upstream airflow. This method of flow induction through the tunnel reduces turbulence and increases flow steadiness, allowing us to have a better comparison to theory. Figure 2 below shows a diagram of the wind tunnel and the direction of induced flow.
The working section of the wind tunnel is where the test article sits. Compressed air is introduced into the flow via the injector block downstream of the working section. Because of the high pressure air that is introduced aft of the working section and directed downstream, this creates a pressure drop in the working section of the wind tunnel and thus inducing air through it. This air moves around the duct, through the flow straightener, through the contraction cone, and then into the working section of the wind tunnel. After passing through the working section, the flow mixes with the compressed air flow and moves around the tunnel again in a cycle. Any excess air that is in the system exits through the outlet filter until the compressed air supply is shut off.
Compressed air is supplied to the wind tunnel by a compressor. The compressor used for this wind tunnel is the Renner Kompressoren RS-PRO 15. This is a 15 kW 220V belt-driven compressor with the capability of compressing air up to 15 bar (about 215 PSI). Air from this compressor then pases through a dryer and a filter system. For this tunnel, we use a Nano D-Series Heatless Compressed Air Dryer. The purpose of this dryer is to remove moisture from the compressed air and filter out contaminants that may otherwise cause unintentional icing and unnecessary wear on the wind tunnel.
Once through the dryer, the air enters 3 receiver tanks. Each of these tanks has a capacity of 500 liters for a total air supply of 1500 liters or just under 400 gallons. The compressor automatically shuts off when the air in these receiver tanks reaches 13 bar. There is a flexible hose that connects these receiver tanks to the the injector block, where compressed air is introduced into the wind tunnel. Also on this line is a compressed air ball valve which can be opened or closed in order to introduce compressed air into the wind tunnel. When open, air is supplied to the wind tunnel from the receiver tanks and when closed, the compressed air supply is stopped. Figure 3 below shows the complete setup of the supersonic wind tunnel in the Low Turbulence Wind Tunnel room in MK 104.
The working section of this wind tunnel is the black section of the wind tunnel aft of the contraction cone seen in Figure 4 below. It is a precision engineered converging-diverging rectangular-section nozzle. That is to say that the the outside of the working section is a rectangle block but the inside of it, where the air actually flows, is a converging-diverging nozzle.
Near the back of the working section is the test section. The test section is where the test article sits. There are two removable glass plates (portals) on either side of the working section that gives access to the test section, allowing us to swap out the test articles. The test articles are held in place via slots in the portals--see above image. The portals are made of glass so that light can be shined through it for Schlieren imaging. Downstream of the working section is the brass injector block which injects compressed air into the wind tunnel.
There is a knob connected to the portals that can be turned to change the angle of attack of the test article. This can be seen in both Figure 4 above and Figure 5 below. On the outer diameter of the portal, there is a gear ring that is connected to a gear on the angle of attack adjustment knob. Therefore when the knob is turned, the gears on the knob turn the gear ring on the portals, thus changing the angle of attack of the test article. This angle of attack is recorded by an angle encoder is read by the data acquisition system.
The bottom of the working section is flat and has 25 static pressure taps along the length of it. The first tap is 19.5mm downstream of the contraction cone and each subsequent tap is 25mm away from the previous tap. A diagram of this is shown in Figure 5 below. Each of these static pressure taps is connected to an electronic transducer that is read by the data acquisition system. Essentially, each of these taps reads pressures along the length of the converging-diverging nozzle.
The top surface of the working section consists of an interchangeable liner. This tunnel is supplied with 3 different liners, which are shown in Figure 6 below. Depending on which liner is inserted into the working section, a different flow speed can be achieved. If the subsonic liner is put into the working section, no supersonic flow can be achieved because the throat area is not large enough to choke the flow to sonic speed (hence the name subsonic liner).
The Mach 1.4 liner can be placed into the working section to achieve flow at a speed of Mach 1.4. This is because the throat area is converged more than the subsonic liner and is reduced enough to choke the flow. It then diverges to an area ratio that gives a flow speed of Mach 1.4. Similarly, the Mach 1.8 liner accelerates the flow to a speed of Mach 1.8 by further reducing the throat area (and thus increasing the diverging area ratio).
The data acquisition system used by this wind tunnel is called VDAS, or Versatile Data Acquisition System. VDAS connects to a software called TecQuipment VDAS that is used to record and export data for this wind tunnel. VDAS reads and records all the relevant data for this wind tunnel and feeds it into the software so we can visualize and export data:
Time
Tunnel Reference Pressure
Supply Pressure
Atmospheric Pressure
Angle of Attack
Up to 32 Static Pressure Taps
This system has the capability of reading up to 32 static pressures from taps labeled 1 through 32. For our purposes, we will only be using taps 1-28. The taps are ordered in a specific way and each tap corresponds to a different pressure measurement.
The Schlieren Apparatus is way to visualize shocks using lenses, mirrors, a lamp, and a screen. Schlieren techniqures were first used in the 19th century to detect faults in glass. The concept of Schlieren imaging for the visualization of shock waves relies on the change in the way light is refracted depending on density variations of the atmosphere in liquids. "Schlieren" is a German word that translates to "streaks."
Parallel light is aimed through the material of interest from one side. The light that leaves the material on the opposite side is focused onto a knife edge. Refracted light falls below or above the knife edge and appears lighter or darker that un-refracted light. This then produces an image of varying light intensity. An example of a Schlieren Image of a shock wave can be seen in Figure 7 below.
The Schlieren Apparatus used for this supersonic wind tunnel is the AF300a Schlieren Apparatus designed by TecQuipment. The system begins with a 100 Watt halogen lamp which acts as the light source. The light from the lamp is focused by passing through a condenser lens. The focused light then travels through an optical slit. The light is then turned 90 degrees using a mirror and goes through an achromatic lens that is flush with one side of the working section portal. The light then passes through the test section and through a second achromatic lens that is flush to the other side of the working section portal. A second mirror turns the light another 90 degrees towards the Schlieren edge. The edge enhances the light refracted image by removing light intensity. After passing through the Schlieren edge, the light then passes through a final lens which then displays the Schlieren image on an imaging screen. A diagram of this process is shown in Figure 8 below.
Figure 9 below shows the setup of our Schlieren apparatus that will be used in this lab. There is a camera placed behind that imaging screen that takes a video of the imaging screen. The video feed from this camera is then displayed on the monitor for easy visualization of the imaging. This system yields the best results when the room lights on that side of the room are switched off--essentially it brightens up the Schlieren image for the camera sin the room is relatively much darker.
A brief description of compressible flows (summarizing important details covered in AE 2010) is given below. It is important to note that inviscid flow will be assumed throughout this discussion. Compressible flow, especially at supersonic speeds, results in some remarkable phenomena. For example, it turns out that an adiabatic flow can increase from subsonic to supersonic speeds if a converging-diverging nozzle is used. If you converge subsonic flow and gradually increase Mach number to sonic speed (M=1) at the throat (minimum area) of a CD nozzle, the flow transitions to supersonic in the diverging section of the CD nozzle!
When a gas flows at Mach numbers greater than M>0.3, the density changes of the gas due to changing Mach number become significant (the flow becomes compressible). Bernoulli's equation, which was derived assuming constant density, is no longer valid for compressible flow. Instead, the relationship between flow characteristics becomes a function of Mach number.
Isentropic flow ocurs when the changes in flow variables are slow and gradual. This results in a reversible process, meaning if the process was reversed, the flow conditions revert to their original values. To visualize this, imagine a flow that is decelerated and then accelerated through a nozzle that gradually diverges and then gradually converges back to its original area. In this process, the flow conditions at the beginning and end of the nozzle are the same. In a reversible process, entropy is conserved, hence the name isentropic flow.
For isentropic compressible flow, a group of equations have been derived to describe the flow. These are shown in Figure 10 below.
If a supersonic flow encounters a solid body, part of the flow must be decelerated to stagnation conditions. Since this necessarily involves a transition from supersonic to subsonic flow, a shock stands ahead of the body. A compression shock is a series of compression Mach waves that coalesce to form one finite shock wave which is comparable to the thickness of the mean free path of the gas involved. Through a shock, an abrupt and instantaneous change in the flow conditions occurs. This process is irreversible and therefore, the entropy of the system increases. Because of this increase in entropy, flow through a shock wave is no longer isentropic and the above equations cannot be used.
The shape of the body determines what shape the shock wave will take. These can be seen in Figure 11 above. In the case of a flat blunt body in supersonic flow, a detached normal shock stands ahead of the body. In the case of a thin wedge, an attached oblique shock stands in front of the body. As a third example, in the case of a thin probe in supersonic flow (such as a pitot-static prove), a small, normal shock stands ahead of the probe tip. A conically shaped bow shock trails from the normal shock. The bow shock weakens as one moves away from its leading edge and, eventually, turns into a Mach wave across which the flow properties no longer change significantly.
Through a shock wave, the Mach number decreases, the static pressure increases, the static temperature increases, and the density increases. Furthermore, the stagnation pressure decreases through a shock and the stagnation temperature stays the same. If the shock is a normal shock, the shock wave decelerates the flow from supersonic to subsonic. A series of equations to calculate flow properties after a normal shock have been derived. These are based on the flow properties before the shock and are shown in Figure 12 below.
Oblique shocks form when compression Mach waves coalesce over a concave curve. When the curve is abrupt (a corner) instead of gradual, the oblique shock wave starts right at the corner. This is shown in figure 13 below.
The equations for the flow through an oblique shock are similar to that of a normal shock. They are heavily dependent on the deflection angle and the Mach number upstream of the deflection. The shock angle of the oblique shock changes if either of these two values are altered. Through an oblique shock, the component of the flow that experiences a shock wave is dependent on the velocity component of the flow that is normal to the shock wave. Because of this, it is possible to have supersonic flow after an oblique shock if the shock angle is small enough! Though it is very easy to derive oblique shock equations from the normal shock equations above, they are posted in Figure 14 below.
When supersonic flow hits a concave corner, it has been shown that an oblique compression shock wave occurs. Similarly, when supersonic flow passes over a convex corner, some sort of expansion shock must occur. An expansion shock in its essence cannot occur the way a compression shock does because it would require a decrease in entropy, which is a violation of the second law of thermodynamics. Instead, a series of expansion Mach waves occurs in place of an expansion shock.
Unlike an oblique shock wave where weaker compression Mach waves coalesce to form one finite oblique shock wave, expansion waves across a gradual convex curve do not coalesce to form one finite expansion wave. This is because while compression mach waves converge across a concave curve, expansion mach waves diverge across a convex curve. Therefore, an expansion shock wave is physically impossible (on top of violating the 2nd law of thermodynamics).
When the flow goes through these series of expansion Mach waves across a convex curve, the flow experiences gradual, reversible changes in flow properties rather than an abrupt change like that in a shock wave. Thus, flow going through an expansion fan is isentropic. The stagnation properties of the fluid are constant, but the static properties change according to the isentropic flow equations described above.
If the convex change is abrupt rather than gradual (i.e. a corner), a group of expansion waves centered at the corner are formed. This is called a Prandtl-Meyer Expansion Fan. See Figure 15 below.
The Mach number increases across an expansion fan and therefore, the static pressure of the flow drops across an expansion fan. The increase in Mach number is dependent on the turning angle of the corner and the speed of the flow before the corner. Figure 16 below shows the expansion fan relations.
In this experiment, a 10 degree wedge is used to gather data and visualize shock waves and expansion fans. Furthermore surface static pressure data is also gathered from this wedge. For your reference, a diagram of the test article is shown in Figure 17 below.
The compressed air supply for this experiment is not infinite and the flow rate through the tunnel is substantial. As such, ensure you are aware of all the data you will need to take for each run of the tunnel and that you know how to take the data ahead of time. The run time for this tunnel stands at about 8 seconds before the pressure in the tanks drops too low and results in an inability to complete the experiment. Ensure you have read the procedure thoroughly and you know what data you will need to take before each run of the supersonic tunnel.
Between runs, you will need to wait about 4-6 minutes for the compressor to charge the tanks. The compressor will charge the tanks to about 10-13 bar. The pressure in the tanks will be too low to accelerate the flow to supersonic speeds when it reaches about 7.5 bar.
Take a tour of the equipment with the TAs. In particular, you will look at and discuss:
The Renner Compressor which feeds the receiver tanks. This compressor is set to kick in whenever the pressure in the receivers drops below 10 bar.
The dryer and receiver tanks. Take a look at the pressure relief valves on the tanks and the tank pressure gauges.
The tubing and valves that supply the receivers and the wind tunnel. Keep a note of which ones are closed and which are open. Try to understand why the tubing is the way it is and why certain valves are open or closed. Ask a TA if necessary.
The wind tunnel itself. Take a look at the tunnel duct, the contraction cone, the working section, and the compressed air injector. Can you figure out how the tunnel works? Why is it shaped the way it is?
The various pressure gauges and pressure taps on the VDAS instrument panel and how they correlate to pressure taps in the working section of the wind tunnel.
The working/test section. Look at how the model is held in place and understand the angle of attack adjustment knob.
The Schlieren Apparatus.
PPE and other safety information about the tunnel.
Preparing the wind tunnel for testing:
Locate the two safety switches and ensure both are switched off. One safety switch is located a little to the left of the supersonic wind tunnel. The second is on the wall to the left of the door above the Yellow Educational Wind Tunnel.
Ensure that the compressor power receptacle is plugged into the black wall outlet extension cable. Ensure that the black wall outlet extension cable is plugged into the wall outlet. These are both twist locks. If either is unplugged, have a TA plug them in for you.
The same outlet is used for the EWT and the Supersonic Tunnel so there is a chance it may not be plugged into the proper equipment.
Flip both safety switches from off to on.
Wait a few moments for the compressor to warm up. It will turn on automatically and run its system checks.
Ensure that the E-Stop on the compressor is not switched on. If it is, turn the E-Stop button clockwise until it pops outward. Have a TA remove the E-Stop error that is present upon startup.
Ensure that the tunnel inlet shut-off valve is closed. This valve is closed when the handle is perpendicular to the pipe. The tunnel shut-off valve is located right below the tunnel pressure gauges on the vertical pipe. DO NOT close the valve on the horizontal pipe!
Turn on the compressor by pressing the green button. The compressor is loud, so use hearing protection if you would like.
When the receivers are empty, the compressor takes about 8-10 minutes to fully charge the receivers. The compressor will automatically shut off when the tanks reach about 11.5-13 bar.
Compare the pressure given by the compressor to the values on the receiver pressure gauges and the supply pressure gauge on the wind tunnel. Are they the same?
Turn on the VDAS panel.
Ensure the orange/yellow extension cord is plugged into the wall and ensure the surge protector is turned on.
Turn on the VDAS by toggling the switch to on. This switch is located on the instrument frame on the left side.
Once the compressor shuts off, zero all the static pressure taps by pressing and holding the “Zero” button on the instrument panel for about 5 seconds.
Setting up the software system for data acquisition:
Open "TecQuipment VDAS" software on the computer from the desktop icon.
If prompted to select the Application, select “AF300 Supersonic Wind Tunnel (Intermittent)” from the menu and click OK. The following screen will open up:
Under Operating Conditions, ensure all 4 boxes are checked so that this data gets exported. Ensure the values read the following:
Liner: Mach 1.8
Atmospheric Pressure (mbar): 1013
Uncheck the checkbox across from "Mach Numbers" so that this data does not get exported. Your software screen should now look EXACTLY like the picture above.
Familiarize yourself with the software. Understand where all the static ports are on the wind tunnel and what all the data on the screen means. Familiarize yourself with where the necessary buttons are for running the software, acquiring data, saving/exporting data, and changing the data acquisition intervals. Have a TA help with explaining this.
When beginning your experiment, one of two airfoil models may already be inserted into the wind tunnel. You may start the experiment with either model in the wind tunnel or with the wind tunnel empty. MAKE SURE you are following the proper procedure for whichever model is currently in the wind tunnel!
For this experiment, you will be running the tunnel without a test article in it. You will be obtaining static pressure data from the static taps along the working section of the converging-diverging nozzle in order to understand the area ratio of the nozzle with respect to the throat of the CD nozzle.
If necessary, have a TA open the test section and remove any test article that may be in the wind tunnel.
Fill the mount holes with the end caps provided.
Leaving the mount holes open creates strange 3D shocks and expansion waves that reflect within the test section leaving us with pressure data that is not representative of the phenomenon we are trying to capture
If the end caps are not available, simply put two models on either side of the wind tunnel such that it is sticking out but the mount holes are filled.
The far side panel is currently an acrylic plate that is slightly thinner than the panel on the front side of the working section. For this far side acylic panel, use the resin printed end cap provided.
A TA can assist with this if needed
Running the Experiment:
In the VDAS software, click the red Play button on the top left corner of the screen. You should now see a live reading of the tunnel reference pressure and the model angle.
In the pop-up window, set the following values:
Interval: 0.5 seconds
Stop after: 10 seconds
DO NOT yet press the play button in the pop-up!
When you are ready to run the experiment, do the following. READ STEPS 4-7 CAREFULLY BEFORE RUNNING THE TUNNEL.
Have one student near the tunnel inlet shut-off valve and also keeping an eye on the tunnel reference pressure gauge
Have one student controlling the software
All students and TAs must now be wearing hearing protection!
In the VDAS pop-up window, hit the red Play button. This will start timed data acquisition.
Wait 1-2 seconds, and then open the tunnel inlet shut-off valve to start flow into the tunnel.
When the tunnel reference pressure gauges reads 7.5-8 bar, close the tunnel inlet shut-off valve.
In the VDAS software, you will see a pop-up notification reading "Timed data capture complete". Click OK.
Navigate to C:\Users\Public\Documents\AE3610Data\<Semester Year>
and select a relevant file name for your spreadsheet. If your section folder does not already exist, create a new one for yourselves.
Upon hitting save, your file will automatically open. Look through the data and make sure it is good and makes sense. Re-run the experiment if necessary.
When beginning your experiment, one of two airfoil models will already be inserted into the wind tunnel. You may start the experiment with either model in the wind tunnel or with the wind tunnel empty. MAKE SURE you are following the proper procedure for whichever model is currently in the wind tunnel!
In this experiment, you will gather static pressure data from the surface of a double wedge airfoil as angle of attack is changed. Furthermore, you will obtain static pressure data along the working section of the converging-diverging nozzle.
Make sure you are taking lab notes.
If necessary, have a TA open the test section and insert the 10 degree double wedge with two pressure tappings test article into the wind tunnel. Make sure the two static taps are facing upwards!
Connect the two static taps to the open tubes near the test section. These will record static tap data to channels 26 and 27 on VDAS. Have a TA help with this if necessary.
The tubes that connect to the static ports have a tendency to get pinched since they are not that long. Make sure both the tubes are not pinched! You will get bad data if they are!
In the VDAS software, click File-->New. This will open a new file for this experiment. If prompted to clear unsaved data, click OK.
Click the red Play button on the top left corner of the screen. You should now see a live reading of the tunnel reference pressure and the model angle.
Rotate the model such that it is at 0 degrees angle of attack
Rotate the model such that it is at -1.1 degrees angle of attack (+/-0.1 degrees is fine)
The encoder is currently zeroed at at -1.1 degrees, hence why we will put it there. However, this is actually 0 degrees in reality!
In the pop-up window, set the following values:
Interval: 0.5 seconds
Stop after: 10 seconds
DO NOT yet press the play button in the pop-up!
When you are ready to run the experiment, do the following. READ STEPS 8-11 CAREFULLY BEFORE RUNNING THE TUNNEL.
Ensure the static port tubes are not pinched
Have one student near the tunnel inlet shut-off valve and also keeping an eye on the tunnel reference pressure gauge
Have one student controlling the software
All students and TAs must now be wearing hearing protection!
In the VDAS pop-up window, hit the red Play button. This will start timed data acquisition.
Wait 1-2 seconds, and then open the tunnel inlet shut-off valve to start flow into the tunnel.
When the tunnel reference pressure gauges reads 7.5-8 bar, close the tunnel inlet shut-off valve.
In the VDAS software, you will see a pop-up notification reading "Timed data capture complete". Click OK.
Repeat steps 5-12 for angles of attack of 2.5, 5, and 10 degrees. You will need to wait for the compressor to recharge between tests.
Keep in mind that this is actually 2.5, 5, and 10 degrees!!!
Navigate to C:\Users\Public\Documents\AE3610Data\<Semester Year>
and select a relevant file name for your spreadsheet. If your section folder does not already exist, create a new one for yourselves.
Upon hitting save, your file will automatically open. Look through the data and make sure it is good and makes sense. There should be 4 data series (one for each angle of attack run) each with 21 data points. Re-run the experiment if necessary.
When beginning your experiment, one of two airfoil models will already be inserted into the wind tunnel. You may start the experiment with either model in the wind tunnel or with the wind tunnel empty. MAKE SURE you are following the proper procedure for whichever model is currently in the wind tunnel!
For this experiment, you will primarily be observing the oblique shock waves and Prandtl-Meyer expansion fans created by a diamond airfoil (double wedge) using the Schlieren apparatus. All data taken during this section will be pictured from either the mounted camera or your camera phone. No pressure data will be obtained.
Make sure you are taking lab notes.
Have a TA remove the two tubes connected to the ports on the wedge.
Setting up the Schlieren Apparatus:
Turn on the Halogen lamp. You can do this by toggling the switch on the VDAS instrument panel.
Turn on the video camera and Monitor.
The camera is turned on by flipping open the camera screen. It is already connected to a power source.
Plug the Micro-HDMI cable into the camera
If the monitor is not already turned on, turn it on by clicking the button on the back of the monitor. You should see an image of the test section and test article from the camera.
Have a TA ensure that the two lenses on either side of the working section are flush to the test section. Do this CAREFULLY.
DO NOT MOVE ANY OTHER LENSES OR MIRRORS ON THE SCHLIEREN TABLE!!!!! These have been PRECISELY set up ahead of time by the lab manager. If anything is moved, have the TA fix it.
Note for TA's: make adjustments to the Schlieren system ONLY if necessary! Call the Head TA or the Lab Manager if you have any issues.
Turn off the room lights so you can see the image better. This light switch is located to the left of the supersonic wind tunnel next to the safety switch that is above the EWT.
Running the Experiment:
Ensure that the VDAS software is running. Press the red play button in the top left if necessary.
Rotate the model such that it is at 0 degrees angle of attack
Rotate the model such that it is at -1.1 degrees angle of attack (+/-0.1 degrees is fine)
The encoder is currently zeroed at at -1.1 degrees, hence why we will put it there. However, this is actually 0 degrees in reality!
When you are ready to run the experiment, do the following. READ STEPS 3-6 CAREFULLY BEFORE RUNNING THE TUNNEL.
Have one student stand near the monitor with their phone camera pointed at the monitor displaying the Schlieren image
Have one student near the tunnel inlet shut-off valve and also keeping an eye on the tunnel reference pressure gauge
All students and TAs must now be wearing hearing protection!
Open the tunnel inlet shut-off valve to start flow into the tunnel. The tunnel will get very loud and you will start to see shock waves on the Schlieren Image.
Take images of the monitor with your phone to capture the oblique shock if you wish.
Alternatively, you may take images with the camera that is mounted and transfer images from the microSD card to your computer using the provided adapter.
This will not be required for your reports
When the tunnel reference pressure gauges reads 7.5-8 bar, close the tunnel inlet shut-off valve.
Wait for the compressor to recharge the tanks.
Repeat steps 2-6 for angles of attack of 2.5, 5, and 10 degrees (again, these are actually 2.5, 5, and 10 degrees!!).
When you are satisfied with all your data,
Turn off the halogen lamp
Turn on the room lights
Turn off the camera by unplugging the micro-HDMI cable and closing the small screen
If you took pictures with the mounted camera, transfer the images to your personal PC using the microSD-->USB adapter provided
Make sure you delete your images from the microSD card after you have transferred them to your PC
Put the microSD card back in the mounted camera
DO NOT leave with the adapter!
This experiment is not necessary for your data report. Do this experiment IF AND ONLY IF (1) you have time remaining at the end of your lab, (2) you have verified that all your data is good and no experiment needs to be redone, and (3) you have no questions for your TAs and nothing to discuss with your group.
For this experiment, you will primarily be observing a detached bow shock created by a blunt body using the Schlieren apparatus. No data or pictures are necessary unless you want to take pictures with your phone.
Have a TA open the test section and insert the blunt body test article into the wind tunnel.
Setting up the Schlieren Apparatus:
Turn on the Halogen lamp. You can do this by toggling the switch on the VDAS instrument panel.
Turn on the video camera and Monitor.
The camera is turned on by flipping open the camera screen. It is already connected to a power source.
Plug the Micro-HDMI cable into the camera
If the monitor is not already turned on, turn it on by clicking the button on the back of the monitor. You should see an image of the test section and test article from the camera.
Have a TA ensure that the two lenses on either side of the working section are flush to the test section. Do this CAREFULLY.
DO NOT MOVE ANY OTHER LENSES OR MIRRORS ON THE SCHLIEREN TABLE!!!!! These have been PRECISELY set up ahead of time by the lab manager. If anything is moved, have the TA fix it.
Note for TA's: make adjustments to the Schlieren system ONLY if necessary! Call the Head TA or the Lab Manager if you have any issues.
Turn off the room lights so you can see the image better. This light switch is located to the left of the supersonic wind tunnel next to the safety switch that is above the EWT.
Running the Experiment:
Ensure that the VDAS software is running. Press the red play button in the top left if necessary.
Rotate the model such that it is at 0 degrees angle of attack (+/-0.1 degrees is fine)
When you are ready to run the experiment, do the following. READ STEPS 3-6 CAREFULLY BEFORE RUNNING THE TUNNEL.
Have one student near the tunnel inlet shut-off valve and also keeping an eye on the tunnel reference pressure gauge
All students and TAs must now be wearing hearing protection!
Open the tunnel inlet shut-off valve to start flow into the tunnel. The tunnel will get very loud and you will start to see shock waves on the Schlieren Image.
Observe the bow shock. Take images if you would like ONLY WITH YOUR PHONE.
When the tunnel reference pressure gauges reads 7.5-8 bar, close the tunnel inlet shut-off valve.
You may wait for the compressor to recharge and run it again if you would like to see the shock again.
Turn off the schlieren system.
Turn off the halogen lamp
Turn on the room lights
Turn off the camera by unplugging the micro-HDMI cable and closing the small screen
Do not start this list if you have any missing or bad data. Go through it with your TAs if you wish.
Hit the red stop button on the compressor panel to turn off the compressor. This button is located right below the green start button. DO NOT HIT THE E-STOP! The compressor will not vent compressed air if you do not hit the proper stop button. It will take about 3 minutes to cool down, vent air, and shut down.
If the E-stop is already engaged, disengage the e-stop, remove the error from the compressor screen, and then start the compressor again
Once the is shown as "On Load" you may hit the red stop button to turn the compressor off properly
Turn off the VDAS by toggling the switch on the instrument panel to off. This switch is different from the Halogen Lamp switch. Make sure BOTH are turn off!
Have a TA upload your data to canvas.
Turn off the TV. DO NOT switch off the surge protector or unplug the orange/yellow extension cable from the wall.
Once the compressor finishes its cool down and turns off, flip both safety switches to off. This will cut power to the compressor.
With hearing protection on, vent the receivers of compressed air by opening the tunnel inlet shut-off valve.
Once the tunnel reference pressure gauge drops to zero, it means the air in the tanks has been discharged. Close the tunnel inlet shut-off valve.
Throw out any used hearing protection and plastic wrappers into the trash bin.
If you are the last section of the day, have a TA empty the water collection tray near the dryer and place it back there.
If the schlieren system is on:
Turn off the halogen lamp
Turn on the room lights
Turn off the camera by unplugging the micro-HDMI cable and closing the small screen
If you took pictures with the mounted camera, transfer the images to your personal PC using the microSD-->USB adapter provided
Make sure you delete your images from the microSD card after you have transferred them to your PC
Put the microSD card back in the mounted camera
DO NOT leave with the adapter!
From the wind tunnel area characterization experiment:
An excel spreadsheet containing 1 data series with 21 data points: the working section converging-diverging nozzle static pressures from the static taps
From the 10 degree double wedge with two pressure tappings experiment:
An excel spreadsheet containing 4 data series each with 21 data points: the working section converging-diverging nozzle static pressures and airfoil surface static pressures at 4 angles of attack (0, 2.5, 5, 10 degrees)
For the wind tunnel area characterization experiment:
Parse your data series to get rid of erroneous data rows (e.g. rows that recorded data before the tunnel was turned on)
From the remaining data, select the row of data you believe is the best row of data
Note: We are leaving this up to you to pick. What is good vs. bad data in this scenario?
Using the row of data selected from step 1.2, calculate the Mach number at each of the 25 stations based on the recorded pressure ratio and given that A* is somewhere between stations 7 and 8
Calculate the area ratio at each of the 25 stations
For the 10 degree double wedge experiment:
Obtain the shock angle, wedge angle, and expansion angle of the airfoil's top surface at 0, 2.5, and 5 degrees angle of attack. Do this using the wedge angle and supersonic flow theory.
Using the Mach number at station 19 as your upstream mach number and using the shock angle information obtained from step 2.1, calculate the mach number and stagnation pressure downstream of the oblique shock.
Using the expansion angle information from step 2.1 and the Mach number and stagnation pressure calculated from step 2.2, calculate the mach number and stagnation pressure downstream of the expansion fan using supersonic flow theory.
Do Steps 2.2 and 2.3 for 0, 2.5, and 5 degrees angle of attack.
For the 10 degree double wedge with two pressure tappings experiment:
Parse your data series to get rid of erroneous data rows (e.g. rows that recorded data before the tunnel was turned on)
From the remaining data, select the row of data you believe is the best row of data for each of the 4 angles of attack
Note: We are leaving this up to you to pick. What is good vs. bad data in this scenario? Your selection criteria to justify this may be a good discussion point!
You should now have 4 rows of data--one for each angle of attack
Calculate the mach number on the top surface of the airfoil downstream of the oblique shock using the stagnation pressure obtained in step 2.2 and the static pressure from the surface tap.
Hint: your pressure ratio for taps 26 and 27 are different from what the software gave you
Calculate the mach number on the top surface of the airfoil downstream of the expansion fan using the stagnation pressure obtained in step 2.3 and the static pressure from the surface tap.
Hint: your pressure ratio for taps 26 and 27 are different from what the software gave you
Do Steps 3.3 and 3.4 for 0, 2.5, and 5 degrees angle of attack.
It is on you to figure out which airfoil surface pressure tap comes first in the flow! Tap 26 could be either after the oblique shock or after the expansion fan depending on how the test article was oriented during your test. Refer to Figure 17 if needed.
For the wind tunnel area characterization experiment at all 25 stations:
A table containing the station number, static pressure, pressure ratio, calculated mach number, and calculated area ratio (A/A*) for the chosen row of data.
A plot of the mach number in the converging-diverging nozzle with the pressure tap distance as the abscissa and the mach number as the ordinate
A plot of the area ratio in the converging-diverging nozzle with the pressure tap distance as the abscissa and the area ratio as the ordinate
For the 10 degree double wedge experiment:
A table containing:
Shock angle on the top surface of the airfoil at 0, 2.5, and 5 degrees angle of attack
Wedge angle on the top surface of the airfoil at 0, 2.5, and 5 degrees angle of attack
Expansion angle on the top surface of the airfoil at 0, 2.5, and 5 degrees angle of attack
Mach number downstream of the shock wave and Mach number downstream of expansion wave at 0, 2.5, and 5 degrees angle of attack calculated from Step 2 of the Data Reduction
If you have them, you may include the four images of the shock at 0, 2.5, 5, and 10 degrees angles of attack. This is not necessary for your report
For the 10 degree double wedge data with two pressure tappings:
For all 27 stations, a table containing the station number, static pressure, and pressure ratio for 0, 2.5, and 5 degrees angles of attack (hint: your pressure ratio for taps 26 and 27 are different from what the software gave you)
A table containing the angle of attack, pressure ratio downstream of the shock, pressure ratio downstream of the expansion wave, Mach number downstream of the shock calculated from pressure ratio, and Mach number downstream of the expansion wave calculated from the pressure ratio
A plot with the angle of attack as the abscissa and the pressure ratio downstream of the shock as the ordinate. On the same figure, plot the pressure ratio downstream of the expansion wave as a function of angle of attack as a different series
A plot with the angle of attack as the abscissa and the Mach number downstream of the shock calculated from pressure ratio as the ordinate (Data Reduction Step 3). On the same figure, plot the Mach number downstream of the shock obtained from the shock and wedge angles (Data Reduction Step 2).
A plot with the angle of attack as the abscissa and the Mach number downstream of the expansion wave calculated from pressure ratio as the ordinate (Data Reduction Step 3). On the same figure, plot the Mach number downstream of the expansion wave obtained from the expansion angle (Data Reduction Step 2).
In this laboratory, students will experimentally investigate the behavior of beams and columns subject to externally applied forces and displacements. First, students will explore the force-deflection and displacement-deflection behavior of a cantilever beam through surface strain measurements. Data from a rectangular rosette strain gauge will be used to determine the state of strain on the surface of a thin rectangular beam. Specifically, rosette strain analysis (strain transformation) theory can be used to determine the principal strain components of a beam under pure bending. Experiments will examine two types of boundary conditions: load-controlled, where prescribed loads are applied at the free end of the beam; and displacement-controlled, where known displacements (measured here with a precision micrometer) are applied to the free end. The measured beam behavior can then be compared to elementary Euler-Bernoulli beam theory predictions to determine Young’s modulus, Poisson’s ratio, and effective beam stiffness. Second, students will explore the buckling and post-buckling behavior of slender columns loaded axially along their length. This part of the experiment includes two tests: a long column and a short column buckling test. Each test will explore the change in critical buckling strength caused by changes in geometry and loading conditions. In addition, students will gain familiarity with buckling analysis techniques such as the Southwell plot.
A beam is a structure whose length is much larger than its other two principal dimensions. In many cases, aircraft structures such as wings and fuselages can be treated as thin-walled beams. Simple beam theory is based on the Euler-Bernoulli assumptions (thus, simple beam theory is also known as Euler-Bernoulli beam theory). Specifically, those assumptions are: 1) the beam cross-section is rigid and does not (significantly) deform under the applied axial or transverse load; 2) the cross-section remains plane after deformation; and 3) the cross-section remains normal to the deformed axis of the beam. These assumptions have been extensively validated for long, thin beams formed from isotropic materials with solid cross-sections, as well as many other beams under “small” deflections
Applying these assumptions allows us to describe the behavior of the beam under load as a one-dimensional function, i.e., solely as a function of the distance along the beam length. For example, the deflection of a beam under an applied load is just a function of the load distribution, the Young’s modulus (E), and the geometry of the cross-section of the beam. The predictions of simple beam theory are available in numerous references, e.g. the text by Bauchau and Craig.
O. Bauchau and J. Craig, Structural Analysis: With Applications to Aerospace Structures, Springer, 2009.
In this lab, we will study a cantilever beam that is subjected to applied loads and applied displacements. A rectangular rosette strain gauge will be used to measure the components of strain on the beam surface near the root in a reference frame that is not aligned with the beam axis. Therefore, it will be necessary to consider a transformation of the components of strain between coordinate frames. If we look at the stresses and strains in a beam subject to transverse load, as shown in Figure 4(a), a good approximation is that only the longitudinal (axial) components of stress exist along the length of the beam. If we use the coordinate frame defined in Figure 4(a), then the axial components of stress correspond to.
Figure 4. Cantilever beam schematic.
Using Hooke's law for an isotropic material, the non-zero observed components of strain, also with respect to the X-Y-Z coordinate frame defined in the figure, should be:
(1)
where E is the elastic modulus and is the Poisson ratio. However, the rectangular rosette strain gauge that is used for these experiments will not be aligned with the X-Y-Z frame. Instead, it will be oriented at an angle. Looking at the beam, we will measure strains in the transformed coordinate system X*-Y*-Z*, shown in Figure 4(a), where we note that the Z axis is shared between the two coordinate frames, and the coordinate system is rotated about the Z axis by an unknown angle θ.
To relate the components of strain in these two coordinate frames, we will use strain transformations. The components of strain in the rotated reference frame X*-Y*-Z* are given as a function of the components of strain in the X-Y-Z frame through
(2)
The strain components in the rotated frame X*-Y*-Z* can be determined with a rectangular strain rosette. For example, Figure 5. shows a strain gauge rosette mounted on the top surface of a beam, with the three gauges separated by 45° from each other.
Figure 5. Photograph of cantilever beam experimental configuration.
Labeling the gauges as A, B, and C as in Figure 4(b), the rosette can thus measure strain in these three directions. If we consider that the frame X*-Y*-Z* is aligned with the rosette such that X* is in the direction, then and are simply given by and . To find we can use the strain gauge in the direction. We consider then a rotation of the X*-Y*-Z* frame to a frame that is aligned in the direction, and using equation (2) yields
Finally, solving for and using and , we can relate the strains measured from the gauges in the strain gauge rosette to the components of strain in the X*-Y*-Z* frame through
(3)
With the strains known in the rotated frame X*-Y*-Z*, we can compute the strains in the reference frame aligned with the beam, that is the X-Y-Z frame. If the misalignment angle θ between the rosette and the beam was known, we could simply use the rotation equations (2). However, it is not necessary to know the misalignment angle since we can recognize that in the beam aligned X-Y-Z frame, all shear strain components are zero, and the strains are principal strains! That is , , and are principal strains under the Euler-Bernoulli beam bending assumption used here. For a two-dimensional state of strain, the principal strains can be computed from the strains in any arbitrary reference frame through
(4)
where and are the principal strains, and then angle from the given reference frame and the principal reference frame is given by
(5)
We note again that the components , , and could be with respect to any Cartesian reference frame.
Figure 6. Mohr's circle for strain.
The strain transformation relations (2) can be visualized using Mohr’s circle as shown in Figure 6, which is a convenient graphical technique for transforming strains (and can also be applied to transforming stresses). All the components of strain in any rotated reference frame lie on Mohr's circle, where a rotation θ in physical space corresponds to a rotation of 2θ on Mohr's circle. In Mohr's circle, the shear strain axis is drawn in the reverse direction such that positive axis points downward. Similar to the equations in (2), Mohr's circle can be used to transform strain components in a X-Y coordinate frame to the corresponding components in an X*-Y* coordinate frame, which is rotated about a common Z-axis by an angle To perform the transformation, first draw the points and . Since all the strain states lie on the circle, straightforward geometry can be used to solve for the strain components in the rotated reference frame. Mohr’s circle also shows us graphically how one can rotate to a particular strain state where the shear strain is zero and there are two principal stresses and .
Slender structures, like beams or columns, that are subject to compressive loads often exhibit elastic instabilities. One type of elastic stability is known as buckling. Aerospace structures are often slender and are subject to compressive loads. Therefore, understanding structural instability in general and buckling in particular is required for aerospace structural design.
Structural instability is a fundamentally nonlinear phenomena, therefore to predict buckling we must model the structural nonlinearity. The most common structural instabilities are due to geometric nonlinearity rather than nonlinear material behavior. Geometric nonlinearity accounts for the effect of deformation within an equilibrium analysis rather than assuming infinitesimal structural deflections.
Figure 1. Euler-Bernoulli beam subject to a compressive load, P.
Consider the buckling of a column loaded by opposing axial loads as shown in Figure 1. We can model this using an extension of Euler-Bernoulli beam theory. Using this theory, the transverse deformation, w(x), of a beam is governed by the equation (1)
(1)
where E and I are the elastic modulus and second moment of area, P is the axial compressive load and q(x) is the distributed load. At first glance, this equation may not appear to be nonlinear, however, the second term is, in fact, a nonlinear term. If we consider a beam that is only subject to a compressive load, with q(x)=0, then the governing equation can be rewritten as
(2)
where .
Note that the governing equation (2) is now an eigenproblem. Using simply supported boundary conditions for the ends of the beam, we find the eigenvalues of (2) are given by equation (3).
(3)
Rearranging this expression gives the following equation for the loads corresponding to these eigenvalues
(4)
We are most interested in the lowest load for which the beam becomes unstable, i.e., the critical load. This corresponds to the case with n = 1, i.e., .
(5)
The critical load based on this theory is often referred to as the Euler buckling load. Using this load, we can compute the axial stress in the beam when it buckles
(6)
where A is the cross-sectional area of the beam. This critical stress is often written in terms of the radius of gyration () of the column about its weak axis, which is defined as . With this definition, the critical buckling load stress and strain from this theory are given by
(7)
The model of Euler buckling is flawed. In particular, the beam is not perfectly straight and the axial compressive load is not applied through the center of the beam. As a result, the beam does not suddenly buckle at but more gradually buckles in a direction determined by the imperfections in the beam. Such an imperfect beam is shown in Figure 2. Even though these imperfections can be quite small, they can have a large effect on the response of the structure. This is called imperfection sensitivity.
Figure 2. Imperfect Euler beam with initial curvature.
A more sophisticated analysis that takes into account an initial imperfection of the beam, where the imperfection causes an initial displacement with the form , results in the following expression for the transverse deflection of the beam at its mid-point () as a function of the Euler buckling load.
(8)
This relationship forms the basis of the Southwell plot. Rearranging this relationship, one can create an expression for the deflection normalized by the load, i.e.,
(9)
A Southwell plot consists of a series of measurements plotted on a graph of δ/P versus δ. The slope of a linear fit to the data then provides and the y-intercept provides a measure of the initial displacement magnitude (see Figure 3a).
Figure 3. Southwell plots for determining the critical load: (a) Southwell plot in original form, (b) Southwell plot for strain. The figures are not drawn to scale, and the slopes on the two graphs are the reciprocal of one another.
In this lab, however, we will only be able to measure the bending strains using the strain gages mounted to the specimen. Therefore, we will modify the classical Southwell plot to use our strain measurements as follows. First, we note that the bending moment at the mid-point of the beam is . With the expression for the bending strain, , the deflection of the beam at the mid-span is
(10)
where we have used as the bending moment. Substitution of this value for δ into (9) yields
(11)
This version of the Southwell plot for strain (see Figure 3b) is the one you will use for data reduction in this lab.
The experiment involves a cantilever beam approximately 12 inches long, which is mounted within a support fixture, see Figure 5. Beam dimensions will be measured with a precision digital caliper. In general, calipers are devices used to measure thicknesses or distances between surfaces, usually having a screw-driven or sliding adjustable piece.
Figure 5. Photograph of cantilever beam experimental configuration.
The set-up allows either an applied load at the free end of the beam (by hanging dead weights) or a prescribed deflection of the beam using a precision micrometer (shown at right side of Figure 5). Precision micrometers typically use a finely threaded screw to accurately measure a linear distance. Turning the outer thimble moves a central rod, called the spindle. A given angular rotation corresponds to a linear distance traveled by the spindle based on the pitch of the screw. The distance the end of the spindle moves can be read from combined scales: a low-resolution length scale on the sleeve and a high-resolution rotating scale on the thimble (see Figure 7). On some micrometers, the readable resolution of the rotating scale is increased by the addition a fine vernier scale on the sleeve.
Figure 7. Micrometer scale.
A rectangular rosette strain gauge rosette is bonded on the top surface near the root of the beam (Figure 5). Each of the three individual grids in the rosette strain gauge will be connected to a dedicated 1/4 bridge circuit within the Vishay 7000 data acquisition system.
All the test subjects are composed of 2024-T3 aluminum. Each column will only be tested in the elastic regime to ensure accurate results with the strain gages and so as not to permanently deform the columns. There are two test specimens: a long and a short column; both have rectangular cross-sections (see Table 1 for dimensions).
Table 1. Dimension of aluminum columns.
When pure axial loads are applied to the nominally perfect columns, we have no advance knowledge of any preferential buckling direction; thus it is difficult to correctly place any lateral displacement gage. As a result, the columns are fitted with two strain gages at mid-span mounted on the top and bottom surfaces of the beam. Assuming that the strain varies linearly through the beam, the axial strain measured at the top () and bottom () surfaces of the beam would be:
(12)
where is the strain from the axial deformation and is the strain due to bending. To find these two strains, we could combine the measured strains to produce
(13)
Due to the fact that the direction of buckling is unknown, the strain reading may either be positive or negative.
Note that the bending strain is also given by
(14)
Measure the beam dimensions
Thickness (h) using the precise electronic caliper
Width (b) using the precise electronic caliper
The length , from the application position of the micrometer to the clamped end of the beam (see Figure 8)
The length , from the point of application of the dead load to the clamped end of the beam (see Figure 8)
Determine the distance between the end of the beam and the rosette center-line.
= 0.8955 in
Note that the accuracy of the distances and are critical to your data reduction
In this experiment you will apply prescribed loads (P) by hanging dead weights from the point labeled “Applied load” in Figure 8, and record the strain readings indicated by each gauge within the rectangular rosette gauge. Follow the steps below:
Prepare the software:
Power on the Vishay 8000 data acquisition unit if your TA has not already done so. There is a switch on the back right of the system.
Open StrainSmart 8000 from the Start Menu
Select New Project on the Top Left. When prompted to save changes to save the new project, Click NO
. When prompted to use the New Project Wizard, Select YES
.
In the Project Settings, you may title the project with your section number (e.g. A12-X). Click NEXT
.
Duplicate/extend the desktop to the large portable monitor
Arrange all relevant windows so everything can be seen optimally (e.g. side-by-side)
Configure the project and scan session:
You should now be on the "Add Sensors" page. From the list on the left, select Rectangular Rosette
. This is the type of strain gauge we will be using for the Cantilever Beam Experiment.
Enter the Gage Factor, Kt value (Transverse Sensitivity) found on the sticker on the beam.
Select a resistance of 120 V from the drop down menu (this is the base resistance of our strain gauge under no excitation). Leave the Default Excitation at 2V.
Click ADD
and when the window closes, click NEXT
.
On the Material Selection page, again click NEXT
.
On the Controller Hardware page, select Connect & Detect Devices
from the list on the left. Click NEXT
.
On the Assignments page, move the cursor over the image of the hardware. You should see the individual ports light up in yellow as the cursor hovers over them. Right click on port 1 and click Add Rectangular Rosette Assignment
.
Verify that Grid 1 is on S1-Ch1, Grid 2 is on S1-Ch2, and Grid 3 is on S1-Ch3. If so, click SAVE
. The first three ports should now have a red box around them.
Click FINISH
.
When prompted to create a scan session, select YES
.
Under session assignments, verify the Rectangular Rosette box is checked and then click NEXT
.
Under scan rate, select 10
and ensure that Time Based Recording is Enabled
under Continuous
Mode. Click NEXT
.
Click NEXT
. Click NEXT
again.
Under relay output, click NEXT
.
The scan session creation is now complete. Click FINISH
.
Click the red DISCONNECTED
button on the bottom left corner of the screen. The red square should turn green and it should now read CONNECTED
. This means the connection between the StrainSmart software and the data acquisition hardware is online and active.
If the system does not connect, ask your TA for help.
Zero the strain gauges:
Ensure the Micrometer is not touching the beam surface
Ensure the hook is on the "Applied Load" part of the beam (the hook is placed in the dimple on the beam)
SKIP THIS STEP FOR THE DISPLACEMENT CONTROLLED EXPERIMENT
Click on Zero Balance
from the choices in the toolbar neat the top of the screen. Click Set Zero
on the bottom of the screen. Ensure there are no errors.
Click on the Shunt Calibration
button near the top of the window and then click Perform Shunt Calibrate
on the bottom of the screen. Ensure there are no errors.
If there was no error message, calibration was successful and click Close. Note that the ethernet connectors in the Vishay box can be a little finicky... if you have trouble with errors being thrown when you zero, or with strain readings drifting in later steps, it is more than likely these connections. To fix this, manually adjust the connectors until the problem stops then take care to not move the connectors afterwards. You may also try different ethernet ports but you will need to change the channel assignment accordingly.
You may have to start a new project if the problem persists. Sometimes errors can be thrown if the settings and numbers you innput do not make sense. Ensure all steps are followed properly.
If problems persist, call the Head TA or Lab Manager for help.
Configure the live display:
Click on Arm and Scan
. From the choices on the left of the window, select Validate Project.
Click Arm
. This gets the software and hardware ready to acquire data.
On the top toolbar, select Open Online Viewer
. Select "Grid 1 Microstrain", "Grid 2 Microstrain", and "Grid 3 Microstrain" from the checkboxes. Ensure everything else is unchecked.
Click Apply Selected Assignments
. Click View Strip Chart
. On the bottom of the screen, ensure the "Auto Y Fit" checkbox is selected. This shows you a graph of Strain vs. Time for each of the 3 strain channels.
Repeat steps 4.3 and 4.4 above but this time, select View Table Data
instead of the Strip Chart. This will open another instance so you can see the actual strain values in real time.
Position both Displays such that you can see them clearly on the TV Monitor
Perform your experiment:
As a group, decide on 5 masses between 100 and 1000 grams, where 1000g must be one of the five.
Don't arbitrarily pick these! Try to discuss with your groupmates and have a reason for selecting the loads you select.
READ STEPS 3-6 BELOW COMPLETELY BEFORE PROCEEDING
When you have your values and justification, click Scan
. You should see your strip chart and table data viewer populated with values.
Ensure that there is no drift in the microstrain values.
If there is, stop the scan, and ensure the strain gauges are properly connected to the Vishay. You will then have to zero the balance again, shunt calibrate, and restart the scan.
For each mass, apply the load at the point labeled “Applied load” in Figure 8 above (there is a dimple on the beam where you apply load), and record the three strains manually.
You will need to hang weights using the hook provided. The hook will rest on the dimple in the beam.
Before recording the strain values, ensure there are no oscillations in the beam.
Discuss all the data you see with your groupmates and TA. Is there anything interesting? Is it what you would expect? Why are you seeing what you are seeing?
Once you have your data, click Stop
.
Remove all masses and the mass application hook and return them to the holder.
When asked if you would like to duplicate the scan session, select YES
. This saves us from redoing all the settings for the following experiment.
In this experiment you will apply prescribed displacements using the precision micrometer used to measure the beam deflection and record the subsequent strain readings indicated by each gauge within the rectangular rosette gauge. Follow these steps:
Because we duplicated the session from above, we will not need to re-enter all the system and strain gauge settings!
Zero the balance and shunt calibrate according to the steps in Step 3 from the Load controlled experiment above.
Perform your experiment:
Familiarize yourself with the micrometer markings and how they relate to displacement
Click Arm and Scan
, and then click Validate Project
.
On your two Online Viewer windows (which should still be open), set one of them to View Table Data
and the other to View Strip Chart
same as you did for the Load Controlled experiment.
Click Scan
in the main window. You should see your strip chart and table data viewer populated with values.
Ensure that there is no drift in the microstrain values.
If there is, stop the scan, and ensure the strain gauges are properly connected to the Vishay. You will then have to zero the balance again, shunt calibrate, and restart the scan.
Zero the micrometer by screwing it down until it just starts to cause a strain, then back off just a touch until it releases the strain
Note down the displacement value on the micrometer. This will be your "zero displacement" value.
Displace the beam to the following positions, measuring and manually recording the subsequent three strains at each position: 0.1", 0.2", 0.3", 0,4", 0.5", 0.6"
Discuss all the data you see with your groupmates and TA. Is there anything interesting? Is it what you would expect? Why are you seeing what you are seeing?
Once you have your data, click Stop
.
When asked if you would like to duplicate the scan session, select NO
. This is your last experiment with this software!
Unload the micrometer so that it is no longer touching the beam
Close down the experiment:
Close any Online Viewer displays that are open.
On the top of the screen, go to the Completed Sessions
tab.
For each of the two sessions, click View Data
. Select Grid 1, Grid 2, and Grid 3 and ensure all the other boxes are unchecked. Click Select
.
Scroll through the Raw Data and look at the Strip Chart. Does this data make sense? Are you satisfied with using this data for your lab report? Verify with your TA if you would like.
If you are satisfied with your data, click the Export Data
button on the top of the main window
Select both sessions from the window to ensure you export both sets of data.
Select "microstrain (Grid 1)", "microstrain (Grid 2)", and "microstrain (Grid 3)" and ensure all the other variables are unchecked.
Export it as an Excel file and have your TA save it to a destination of their choosing.
Click Export
.
Open the Excel file and verify all data was exported properly.
Go to the Configuration
tab near the top of the main window. Click EXIT
. When prompted to archive the project, select NO
.
Power down the Vishay data acquisition unit by flipping the switch on the back of the box.
Have your TA upload your data to Canvas.
Turn on the load frame using the power button on the right hand side of the base
Open Bluehill Universal and Click "Test"
Under "New Sample" select "Browse Methods"
Navigate to C:\Users\Public\Documents\Instron\Bluehill Universal\Templates\AE_Labs\AE3610\Bending Buckling
Open the file titled AE_3610Buckling.im_tens
and click OK
Familiarize yourself with the software, shown below. This experiment will be performed manually using the jog function and the fine position dial on the Instron control panel. Look for the following:
Displacement (as measured by the load frame's motor controller)
Force (as measured by the load cell on the crosshead)
Strain Gauge Adapter Module A and B (strain measured by the strain gauges wired into strain gauge adapters A and B)
The Zero Displacement
and Balance All
buttons enable you to balance (aka zero) the sensors, which will not be balanced when you first open Bluehill Universal
Locate and install the long column:
With nothing installed in the load frame, hit Balance All
in Bluehill to zero the load cell.
Jog the load frame manually, using the jog buttons and fine adjustment wheel on the control panel, until the gap between the v-grooves is just larger than the long column.
Sometimes the Jog Down button doesn't work; if this happens, have a TA jog the crosshead down using a quick compression method. Do not do this yourselves! the TAs are trained to handle this scenario!
With your hands in a safe position so they can't get pinched, use the fine adjustment jog wheel to jog the top frame down until the column is held in place. Find the exact correct position by observing the Force
reading in Bluehill - it will start to shoot up once the upper v-groove makes contact. At this point, back off slowly until the column is just unloaded. Ensure the column doesnt slip out of the grooves.
Install and calibrate the strain gauges:
Full scale: 5%
Gauge Length: 1 mm
Calibration Type: Automatic
Calibration point--Values below as of 10/17/2023 (Calculated by 10/Gauge Factor)
Long Column Calibration Point = 4.93%
Short Column Calibration Point = Written on the Column itself
The two gauges have different gauge factors for each gauge since the strain gauges are from two different batches
Since the gauge factors are different, the two gauges have different calibration point values. These calibration point values are written on the column itself for each side
Make sure you know which gauge is plugged into Transducer A/1 and Transducer B/2 so you can enter the appropriate value for each transducer
Offset: 0%
Click Calibrate
and then Click OK
If the calibration fails, check the wiring at the strain gauge completion box - usually it fails when there is an open circuit which can come from a faulty connection or damaged strain gauge
Note: If the strain gauge transducer icon turns gray, it means that the transducer did not calibrate properly
Once calibrated, click Balance
and then click Close
Close the system details window
Click Balance All
in the home screen
Observe the strain gauge readings in Bluehill - if they are erratic or drifting, you possibly have a bad electrical connection. In this case, adjust the wires and re-balance as necessary until you have the required behavior.
Without touching the strain gauges directly, perform a sanity check on the setup by forcing the column to the left and right with your hands. Verify what you expect to happen against what you observe on the live Bluehill strain readings.
LOAD STAGNATION--Read this step carefully before continuing!
In the steps below, you will load and unload the column to 9 separate values.
Sometimes, the load stagnates despite increasing crosshead displacement--this means that even though you increase displacement using the fine adjustment wheel and the column continues to buckle, the load reading stays constant.
To prevent damage to the column and strain gauges, please following these steps:
Continuously ensure that the applied load is increasing as the fine adjustment wheel increases crosshead displacement
Keep an eye on the columns to make sure it is not visibly buckling too much. These columns can and do yield if too much load is applied and the columns then become useless! It can also cause the strain gauges to fail or pop off!
If the load does stagnate during your experiment, follow these steps:
Unload the column completely
Visually inspect the column to make sure there is no plastic deformation
Hit Balance All to zero everything
Re-insert the column into the grooves
Hit Balance All again
Continue the experiment from where you left off
A TA will help with this process.
Commence the loading experiment:
You will load the column to five values of load between 100lbf and 900lbf (inclusive)
100 lbf
300 lbf
500 lbf
700 lbf
900 lbf
Have one student open a blank spreadsheet and be ready to record the data
Using the fine adjustment jog wheel, load the column to each of the five load values. At each point:
Visually observe the column for buckling
Log the exact load and both strain gauge readings into your spreadsheet
Commence the unloading experiment:
You will unload the column to four different values of load and one same value of load between 100lbf and 900lbf (inclusive)
800 lbf
600 lbf
500 lbf
400 lbf
200 lbf
Using the fine adjustment jog wheel, unload the column to each of the five load values. At each point, log the load and both strain gauge readings into your spreadsheet
Verify your data:
In your spreadsheet, add columns for, and calculate, axial and bending strain from your recorded strain gauge data
Make a quick scatter plot of bending strain to sanity check the data with the help of the TAs. If anything looks off, now is the time to repeat the experiment.
Unhook the strain gauges from the strain gauge adapter module. Make sure to save the jumper wires!
Remove the long column by manually adjusting the fine adjustment wheel until it can be freed.
REMEMBER TO KEEP AN EYE OUT FOR LOAD STAGNATION
Repeat all steps from the long column test, but this time use loading/unloading values in the range of 500lbf to 1300lbf (inclusive).
Use these loads for loading:
500 lbf
700 lbf
900 lbf
1100 lbf
1300 lbf
Use these loads for unloading:
1200 lbf
1000 lbf
900 lbf
800 lbf
600 lbf
Verify your data! Redo the experiment if anything looks off.
As before, unhook the strain gauge from the strain gauge adapter module and remove the short column from the load frame by manually adjusting the fine adjustment wheel until it can be freed.
Return the jumper wires to the green/black toolbox on the table!
Close Bluehill. If promped to save the sample, select "No". Turn off the Computer.
If you are the 8:25am section or the last section of the day, power down the Instron. Otherwise, keep it on.
*These instructions are for the linear strain gauges currently installed on the columns, which are 2-wire Omega KFH-3-120-C1-11L1M2R strain gauges (as of February 2022). Consult the strain gauge adapter module manual for new values/wiring instructions if these ever change.
Cantilever beam dimensions and locations of rosette, micrometer and load along beam.
Strain readings for known loads from the cantilever experiment.
Strain readings for known deflections from the cantilever experiment.
Axial loads and bending strains for loading and unloading runs for long column buckling.
Axial loads and bending strains for loading and unloading runs for short column buckling.
Using the rectangular rosette strain gauge data, determine the strains , , and and that resulted from each of the applied loads.
Based on , , and estimate for each of the applied loads: the maximum and minimum principal strains ( and ).
Determine the slope of the best fit linear relationship between and P and use it, along with cantilever beam theory for the relationship between the axial strain and P at the free end of the beam, to estimate the elastic modulus E in MPa of the cantilever beam material. This relationship will also depend on the geometric quantities and
Using the rectangular rosette strain gauge data, determine the strains , , and that resulted from each of the applied deflections.
Based on , , and , estimate for each applied deflection: the maximum and minimum principal strains ( and ).
Determine the slope of the best fit linear relationship between and , and use it, along with the relationship betweenand the beam deflection from cantilever beam theory to estimate the “effective bending spring stiffness” (in N/m) of the cantilever beam. To accomplish this, you will also need to use your estimate of from the load-controlled data
Using a Southwell plot for each of the columns, determine the column’s buckling load and buckling (axial) stress.
Calculate the theoretical buckling load and stress for each column based on column theory.
Table of measured rosette strains and inferred local strains for each of the applied loads.
Table of measured rosette strains and inferred local strains for each of the applied deflections.
A combined plot of and as a function of the applied load (in Newtons).
A combined plot of and as a function of the applied deflection (in meters).
Plot of as a function of (in Newtons), including the best-fit line.
Plot of as function of (in meters), including the best-fit line.
Table of linear proportionality constants for vs. P and vs. , and inferred beam elastic modulus and effective bending stiffness.
Table of measured strains for each of the applied (measured) axial loads for long column.
Table of measured strains for each of the applied (measured) axial loads for short column.
Individual Southwell plots for each of the columns.
Table of experimentally determined and predicted buckling loads for each column.
In this lab, you will learn the basic principles of laser Doppler velocimetry – which is a standard technique for measuring instantaneous, spatially resolved flow velocities. You will then use a laser Doppler velocimeter to explore the flowfield of a jet in a coflow. The jet is produced by the exhaust of a round pipe centered in a small, low-speed wind tunnel.
Rapid fluctuations in velocity occur in most flows of practical interest. Examples are: the boundary layer above wings and fuselage surfaces, the wakes behind obstacles in flows, jets from the nozzles of rockets and gas-turbine engines, and flows inside engine components. These fluctuations have profound effects on such things as drag, surface shear stress, boundary layer separation, mixing between fuel and air in engines, and vibrations of turbomachines and control surfaces. To understand such phenomena, we must be able to measure velocity fluctuations accurately.
There are several difficulties in making such measurements. First, we must find a method that has a measurable change in output for small changes in velocity (good signal sensitivity or resolution), and which will work in the unsteady environment of a fluctuating flowfield (robust). Second, the device must respond faithfully, and perhaps without any time lag, to rapid fluctuations (good temporal resolution or good high frequency response). Third, the device must respond only to the velocity in a very small, and precisely known, region (good spatial resolution).
While the Pitot probe used in earlier experiments can measure mean velocities, it is not suited to measuring fluctuation velocities. There is another well-known probe, the hot wire (also known as a constant temperature anemometer) that is suited to this task (under certain limitations). Its small, almost “microscopic” size gives it good spatial resolution and high frequency response. In addition, it has good signal sensitivity to velocity in constant temperature gas flows. The sensing element is a long circular cylinder, typically a tungsten or platinum wire of diameter in the range of 5 to 20 µm (2-8 ×inches). Slightly more rugged are hot-film sensors, which use glass rods, ~50 µm in diameter, with platinum films, typically 10 Å thick (Å = Angstrom = m), coated on the surface. In either case, the wire or the glass rod is fixed to two gold-plated steel needles which serve as the electric contacts to the sensor. The wire or film is kept heated by an electric current, to temperatures of 200 to 300 °C (and sometimes up to 800 °C). When air flows over the wire, energy in the form of heat is carried away by the much colder air stream (forced convection) because of the temperature difference between the (hot) sensor and (cold) air.
In many cases, however, probe-based approaches should be avoided so that the measurement not interfere with the flowfield (nonintrusive), else it change the quantity being measured. In addition to altering the flowfield, there may be several other reasons for using a nonintrusive technique. First, it may be physically difficult to design a probe that can survive in the flow to be measured. For example, consider the problem of measuring velocity inside the passages between the blades of a jet engine compressor while the engine is operating. Also it can be useful when several properties are varying simultaneously, as in a turbulent flame (e.g., inside an engine combustion chamber). For example, the velocity, pressure, and temperature may all be changing from one instant to the next. Most probes, such as Pitot tubes and hot-wires, respond to changes in several of these properties, and so it may be quite difficult to separate out the part due to velocity alone. Still another problem arises in two-phase flows (e.g., not all gas): there may be liquid droplets or solid particles that can clog or even instantly destroy sensitive probes. Examples of such applications include injection of liquid propellants or fuels sprays in rocket and aircraft propulsion systems, and water droplets and ice particles flowing around wings.
The laser Doppler velocimeter (LDV, also known as a laser Doppler anemometer - LDA) is a non-intrusive measurement device that is sensitive only to velocity. Thus, it is one of just a few techniques that can be used in complex flowfields to measure velocity. This is why, despite the greater cost and difficulty usually involved in making measurements with this technique, laser velocimeters are highly regarded instruments.
When electromagnetic waves are scattered from moving objects, the scattered light has a frequency which is different from that of the incident waves. This is similar to the phenomenon that we have all observed: The horn of an oncoming car changes pitch suddenly as it passes us. The difference in frequency is known as the Doppler shift, and is proportional to the relative velocity between the moving object and the observer (or receiver). This principle is also used in radar and laser speed detection.
A laser is a source of radiation that generally has most of the following properties; it is: 1) well-collimated, which means the beam will not diverge or change in size much; 2) highly coherent, meaning its electromagnetic wavefronts are all in phase; and 3) nearly monochromatic, which means that the radiation is composed of a very narrow range of wavelengths. Thus, it is ideal for making measurements of Doppler shift; you can precisely and easily calculate what the shift should be for a given velocity. Unfortunately, the frequency of light is very high, and the Doppler shift caused by the kinds of velocity that we encounter is extremely small by comparison. So, if one were to shine a laser beam at a moving object, measure the incident light frequency and the scattered frequency, and try to find the difference, one would probably find that one's measuring instrument cannot make out the slight difference between these frequencies. The solution to this problem is to find the difference in Doppler shift caused by two beams, hitting the same object at slightly different angles. If one tries to measure only this difference, one can get good accuracy. This is the principle of the dual-beam interferometer-type LDV that is widely used today (see Figure 1).
Figure 1. Schematic of a basic 1-component LDV system in forward-scattering mode (plan view).
When two laser beams of the same frequency cross in a small region, stationary interference fringes are formed, as shown in Figure 2. This is because light behaves as a kind of wave, with an amplitude and a phase. At some points, the two beams are in phase and the amplitudes add together; there the light’s amplitude is large. At other points, the amplitudes cancel each other out (out-of-phase), and you get nothing at all. Thus, fringes are formed, with each bright fringe being a thin disc inside which the light intensity is large. The bright fringes are separated by dark fringes, where there is hardly any light intensity at all.
Figure 2. Plan view of fringes formed when two laser beams cross (particle motion across fringes will yield vertical velocity in plane of paper). Note that the region with fringes looks ellipsoidal in shape
As seen from Figures 2 and 3, the region where the fringes exist is an ellipsoid of revolution. Consider a small solid or liquid particle moving perpendicular to the fringes (as shown in the figure). At the same time, imagine that a highly sensitive light detector is looking straight at the light being scattered from the ellipsoid. As that particle enters the first fringe, it starts scattering more light. The detector sees the received light intensity increasing. The intensity reaches a maximum when the particle is centered on the fringe, and decreases thereafter. As the particle reaches the center of the first dark fringe, the intensity sinks to a minimum, and the slowly starts increasing again as the particle approaches the next bright fringe.
Figure 3. Sinusoidal variation in the intensity of scattered light, as a particle crosses the fringes (d=fringe spacing; U=velocity component perpendicular to the fringes; Δt=time between scattering peaks or 1/Δt = Doppler frequency).
As the particle crosses the ellipsoidal region (the measurement volume of the LDV), the detector sees the light intensity oscillating in a sinusoidal fashion; this is known as a signal burst (as seen in Figure 3). If the particle moves faster, the same number of fringes will be crossed in less time, so that the frequency of the sinusoidal signal will increase. Thus, the frequency of the signal received at the detector is proportional to the speed of the particle along the direction perpendicular to the fringes. Note that it does not matter, as far as the frequency is concerned, how fast the particle is moving in other directions. Only the speed at which it crosses the fringes is important. Thus, the dual beam LDV is sensitive only to the component of velocity that is perpendicular to the bisector of the two laser beams, and in the plane of the beams.
Let us try to quantify this relationship. The frequency of the detected signal (which can be rigorously shown to be exactly equal to the Doppler shift) must be directly proportional to the velocity component, and inversely proportional to the fringe spacing (the distance between the middles of two successive bright fringes). The fringe spacing, d, for two beams of wavelength λ, intersecting at a half-angle θ, is given by:
(1)
Thus, the Doppler shift frequency is simply,
(2)
where U is the velocity component perpendicular to the fringes.
As an example, consider an Argon-ion violet laser with a wavelength of 476.5 nm (not the one in our LDV system). If particles in an air flow are moving at 10.0 m/sec and the half-angle between the laser beams is 6.0°, then the fringe spacing is 2.28 μm, and the Doppler shift is 4.39 MHz (a frequency of 1 MHz means that the time between two successive peaks of the signal issec).
Notice that in the above calculation, the only parameter that depended on the flow was the flow velocity; the other parameters (laser wavelength and half-angle) depend on the instrument. Thus, the LDV requires no calibration, unlike many other measurement techniques. The half-angle (θ) between the beams is determined using simple geometry knowing the optical setup, i.e., by the spacing (D) between the two beams when they come out of the front focusing lens of the LDV, and the focal length (f) of the lens, which is the distance from the middle of the lens to the focal point where the beams cross (see Figure 1).
While the simple dual-beam LDV system (such as that shown in Figure 1) is able to isolate and measure a single component of the flow velocity, it can not differentiate between positive and negative velocities. If the flow were to change direction, e.g., from forward to backward, a particle moving through the fringes with the same velocity magnitude would produce the same Doppler shift frequencyfor either direction of motion. To overcome this, one of the two laser beams in a dual-beam LDV system can be passed through an acousto-optical modulator* (called a Bragg cell) that can shift the beam’s frequency slightly (e.g., by 10s of MHz).
*In the acousto-optic modulator, the laser beam interacts with sound waves generated (typically by piezo-electric transducers) in an optically transparent material. The laser beam is diffracted off the index-of-refraction pattern caused by the pressure waves in the material.
When the two laser beams intersect, the fringes that are created are no longer stationary in space. The fringe pattern (i.e., the locations of the peaks and valleys) moves at a rate determined by this frequency shift. Now the detector will see a signal with an amplitude that oscillates at a frequencythat is different than . Assuming that the fringe pattern moves in the “forward” or “positive” flow direction, the signal frequency is given by
(3)
For example reconsider the particle describe above, which was moving at 10.0 m/s through a fringe pattern with a spacing of 2.28 μm (= 4.39 MHz). If the second laser beam was shifted by a Bragg cell with frequency 10.0 MHz, causing the fringe pattern to move at 10 MHz in the negative direction of the flow, then the light scattered by the particle as it moved through the measurement volume would have an amplitude that oscillated at a frequency of 10+(+4.39)=14.39 MHz if the particle was moving in the forward direction, or 5.61 MHz if it was going backward.
The LDV system for this lab uses a 300 mW green laser, fiber-coupled to a transceiver unit (see Figure 4). The light coming out of the laser goes to a fiber-coupler unit that first uses a beamsplitter to separate the original beam into two beams. One of these beams is then passed through a Bragg cell that shifts the frequency by 40 MHz. The two beams are then sent to the transceiver unit through fiber optic cables. The laser beams pass through a lens in the transceiver, and are focused down, so that they cross at the focus of the lens and form the measurement volume.
Figure 4. Schematic of lab LDV system with combined transmitter and receiver (transceiver) used in back-scattering mode.
The back-scattered light from the focal volume is collected by a lens located in the transceiver, and focused in front of a photomultiplier, a light sensitive detector that can detect even very low light levels. The photomultiplier, when activated by a high voltage across it, converts incident light (photons) to an electric current (electrons) with a significant gain. Typically, a single input photon striking the photomultiplier can result in a large number (e.g., ) output electrons. This electron current, when passed across a high resistance, produces a voltage that we measure (see Fig. 4 in the Unsteady Combustion Lab). The photomultiplier output goes to the Photodetector Module (PDM), which also controls the photomultiplier input voltage (and thus the detector’s gain).
Since the receiving optics are located on the same side of the tunnel as the transmitting optics, our LDV uses the backward scatter (or backscatter) configuration. It turns out that when the diameter of a scattering particle is about the size of the wavelength of the light, most of the light scattered by the particle goes right past the particle, for example a typical value would be 90% of the light in this forward scattering direction. Only about 1% of the light comes straight back along the path of the incident beams, in the backscatter direction. On the other hand, our usual experience of looking in mirrors tells us that, when the size of the scattering object (say, a mirror) is much larger than the wavelength of the light, the light will not go around the object, but gets reflected right back. Thus larger particles will scatter a larger fraction of the incident light in the backward direction.
The output from the PDM goes to the FSA (Frequency and Size Analyzer) digital signal processor. The burst signal is first sent through an 80 MHz low pass filter, then downmixed with a user-chosen frequency; essentially the user-chosen frequency is subtracted from the signal. You can use this to subtract part of the shift in the signal frequency that was added by the Bragg cell. For the example given above (= 4.39 MHz, 10 MHz Bragg shift), a downmix frequency of 5 MHz would convert the 14.39 MHz burst frequency to 9.39 MHz.
Selecting a downmix frequency of 40 MHz will eliminate all the 40 MHz Bragg Cell Shift, so only the Doppler frequency is left, which is proportional only to the particle velocity. Selecting 39 means 39 MHz is subtracted from the 40 MHz Bragg cell shift, leaving 1 MHz still added onto the Doppler frequency. This will allow flow reversals up to 1 MHz to be measured. With a 5 µm fringe spacing, this would correspond to flow reversals up to 5 m/s. Entering 36 means 36 MHz is subtracted from the 40 MHz Bragg cell shift, leaving 4 MHz still added onto the Doppler frequency. This would correspond to 20 m/s (assuming 5 µm fringe spacing), in terms of the maximum flow reversal velocity.
After the downmix subtraction, the signal is sent through a bandpass filter to remove content at frequencies that are nowhere near the expected frequencies based on the estimated flow velocities; this is required because there are many possible sources of spurious electrical signals (“noise”). If what remains is the true (Doppler shift) signal caused by the particle crossing the fringes, the remaining signal would ideally just have one frequency. After being filtered, the digital signal processing units digitize the signal and determine the frequency embedded in the burst (similar to what could be achieved with a discrete Fourier transform). This part of the FSA electronics is called the burst correlator.
Normally the first parameter to select is the Band Pass Filter setting. If you know the range of frequencies that will be in the flow, you can select the setting directly. Remember that the signal frequency (without frequency shift) is equal to the particle velocity divided by the fringe spacing. To estimate the frequency of the signal input to the signal processor, add the effective frequency shift (equal to 40 MHz minus the downmix frequency).
The system also includes a threshold for detection. Unless the burst signal has an amplitude of at least some minimum value, say 30 mV, the electronics can reject it. For example if the signal is too weak and buried in noise, no useful velocity measurements can be made. This fact puts the burden entirely on the operator of the system to ensure that the signal going into the correlator is of good quality, (good visibility). In other words, if you can not see a clear Doppler signal on an oscilloscope, do not expect the burst correlator to see one either. Of course, if you amplify a noisy signal enough, the unit will faithfully find some frequencies in the noise signal, and may output useless data. The frequency counted may be that of the nearby radio station instead of the Doppler shift, and you may get supersonic flow velocities from your low-speed flow!
Note that a valid signal results only when a solid or liquid particle crosses the measurement volume. The velocity measured is actually of these particles, not of the air itself! Thus, care must be used in ensuring that the particles present in the flow are those that are so small that they move at essentially the same speed as the flow. In other words, they must have very small mass, and comparatively low density. If they move at speeds different from the local air flow velocity, the drag force exerted on them must be far greater than the inertia due to their mass, so that they are again very rapidly accelerated or decelerated to the local air velocity. You can easily convince yourself that as the diameter of a particle gets smaller, this becomes a better approximation, since the mass of a sphere is proportional to the cube of the diameter, while the surface area is proportional to the square of the diameter.
Usually, in a laboratory, the air is supposed to be clean and free of dust particles. Hence, the flow must be artificially seeded with particles of the right diameter range. This is done here by means of an atomizer, which works essentially like a paint spray can. A high-pressure air supply is expanded through a small orifice or nozzle to high speed. On the way out, it hits a thin stream of liquid (mineral oil) and shatters it into tiny droplets. This flow is then rapidly decelerated and turned around a sharp corner, and forced to go upwards. All the large droplets hit the walls (low drag, high inertia - they can not turn quickly enough) and drip back down into the oil reservoir. The remaining droplets, with diameters in the range 1-10 μm, go with the air stream and mix with the main flow. If done properly, a uniformly seeded flowfield can be created.
Our LDV system uses an in-house fabricated wind tunnel, seeding system, and manual traverse, with an off-the-shelf LDV system made by TSI Inc. The three figures below show our setup, its primary constituent parts, and a view of the test section when the laser transceiver is energized.
Turbulent flows are not only unsteady, but chaotic in character.*
*For background material on turbulent flows, you may wish to read Anderson’s Fundamentals of Aerodynamics, 1984, Chap. 16, or Kuethe and Chow, Foundations of Aerodynamics, 4th Ed., Sections 18.1 - 18.3.
In analysis of many turbulent flows where the fluctuations in velocity are small compared to the mean velocity, we decompose the instantaneous velocity u(t) as follows
(4)
where is the mean velocity and u′ is the fluctuating component (with u′<<). These may be seen from Figure 5, which illustrates how the velocity at some point would change with time in a turbulent flow.
Figure 5. Typical velocity history at a single point in a turbulent flow.
Note that the time average of the fluctuating component, , should be zero since the fluctuations are random and the u′ at any instant is as likely to be positive as negative. However, the variance is not zero, and gives us a measure of the intensity of the fluctuations. The positive square root of the variance is called the root-mean-square, or RMS for short. The variance of the velocity enters directly into the time-averaged Navier-Stokes equation (as described in AE 3030). This quantity will be calculated and used here.
The velocity may be expected to vary from one data point (i.e., one particle) to the next. The LDV system collects, say, the first 5,000 points to be measured, converts the time values to velocity, and then sorts these values into “bins” or velocity intervals. Thus, if the highest velocity that you expect is 20 m/s, and the lowest is 10 m/s, and you divide this range into 100, then all the data points which show velocity values between 10.0 and 10.1 m/s will go into the first “bin”, all those between 19.90 and 20.0 m/s will go into bin #100, and everything else will go somewhere in between. The mean velocity can be calculated from the binned data with the formula
(5)
whereis the number of samples in the i’th bin, and is the center velocity of that bin. The root-mean square (or the square-root of the variance) is calculated as:
(6)
These results are usually displayed on the computer, along with a graphical display of the distribution of data points over the velocity intervals. Such a graph is called a histogram. The number of points in each interval has been divided by the total number of points collected. If it were a continuous function, instead of being a series of discreet values for a finite number of intervals, it would be called a probability density function (PDF), whose integral, over the entire range of velocities, should thus be equal to unity.
Usually when this procedure is used, at leastpoints are needed to yield a good result for the average velocity, and at least-for the RMS. This is if you collect data for a sufficiently long time, so that you have seen all the likely kinds of fluctuations that occur. However, if the flowfield is fluctuating slowly, the average calculated from just one set of data as above may be insufficient. Instead, you may have to average over many data sets, each taken at a different time. For example, suppose there are 6000 data points arriving every second, and you collect the first 3,000. This should take only 0.5 seconds. Suppose now that the flow itself is fluctuating so that the magnitude of velocity goes up and down again once in 2 seconds. Your calculated mean velocity will not be the true mean of the flow. To find the true mean, you will have to average many such data sets. Note from Eq. (6) that you cannot just average the rms values to get a better estimate for the rms; this would have to be done inside the summation sign.
One feature of the laser velocimeter to bear in mind is that one must have a good idea of the flowfield before attempting to make measurements using it. Very precise measurements can be made by a careful and alert experimentalist, but not without some consideration of the flowfield. Figure 6 shows a schematic of the flow setup in our laboratory. A fan downstream of the tunnel diffuser creates a pressure difference that induces the tunnel freestream, and forces it out of the laboratory. A circular pipe brings air output from a centrifugal blower to the upstream edge of the wind tunnel test section. The flow comes out of the pipe as a jet, with a velocity that can be significantly higher than that of the test section freestream. Both the test section freestream and the flow inside the pipe are seeded with atomized mineral oil.
Figure 6. Schematic of jet flow in the wind tunnel
If the jet flow leaves the pipe with a high velocity, the effect of viscosity at the edges of the jet causes the surrounding test section flow to accelerate, and the jet flow to decelerate. In fact, the flow at the edges of the jet is forced outward as it is “pulled” by the slower-moving test section air. Thus, some of the jet flow “rolls up” into vortex rings around the jet, so that some air which originated in the test section flow goes into the jet, and vice versa. Thus, there is a mass transfer and momentum transfer across the jet boundaries. Things rapidly become confused and unsteady at the jet edges, so that the velocity fluctuates. Not too far downstream of where the jet exits the central pipe, the fluctuations are largest inside the shear layer at the edges of the jet. Near the center of the jet, there is still a region where the fluctuations have not reached. This is called the potential core, and the fluctuation level should be quite low here. The shear layer around the jet grows quickly so that the shear layers from opposites sides merge, and the potential core rapidly disappears as the jet goes downstream. At the same time, the mass and momentum transfer force the time-averaged velocity profile across the jet to become smoother and to gradually disappear, as the jet merges with the test section flow. You will acquire velocity data along a vertical line extending from somewhere below the top wall to a region below the centerline of the jet. This line is nominally near the horizontal center of the jet and nearer to the end of the test section.
The specifications for our LDV system and software configuration are as follows:
Laser wavelength, : 514.5 nm (Argon-Ion green)
Beam separation, : 50 mm
Focal length, : 362.6 mm
Approximate range of measured velocities, : -2 to 15 m/s
Bragg cell frequency: 40 MHz
Burst threshold: 30 mV
Calculate the following parameters:
Beam intersection half-angle,
Fringe spacing,
Doppler frequency, , at each end of the velocity range (assuming no Bragg)
Assuming no bandpass filter is implemented, what is the absolute minimum downmix frequency needed to enable the required negative velocities to be measured?
Given a downmix frequency and an ideal bandpass filter that operates between and Hz, write a closed-form expression for the minimum and maximum observable velocities of the final system
Using your answer from Q5, select an integer-number downmix frequency to input into FlowSizer, and select a band pass filter from the available options in "LDV Controls" (see pic below for precise options). Provide the minimum and maximum observable velocities using these settings.
Hint: everything you need to answer these questions is in this manual, but you may also get some better insight from the LDV manual (linked below), and the FlowSizer auto min and max velocity calculation
Please note the following safety precautions very carefully:
Your eyes, and your friends' and instructors' eyes, can get hurt by lasers. Take no chances or shortcuts: laser beams travel a lot faster than you can react! To ensure safety, no one who has not read the instructions thoroughly can be allowed near the experiment. Anyone disobeying safety rules will be ordered out of the lab immediately.
Wear laser eye protection at all times once the laser has been powered on!
Remove all rings and watches when working in the vicinity of lasers.
Do not go to the back side of the test section unless directly supervised by a TA.
Do not lean over the laser beam, or stick your head between the laser transceiver and the test section when the laser is on.
Do not touch the laser or fiber optic coupler, whether the laser is on or off. If they go out of alignment, the TA will realign the system. Also, do not touch the transceiver unless instructed to.
Be very careful where you move the traverse: you must know clearly where all the reflections are (if any)!
The TAs will now give you an overview of the equipment and lead you through configuration of the experimental hardware and software.
We will now go through a sample experiment to demonstrate operation of the system and how to acquire data:
Verify the traverse is set such that the transceiver probe is aligned in the center of the jet and the direction of the probe is in the direction of flow (the yellow warning label is on the left when viewing the probe from behind, and angle dial is on 90)
Open both seeding valves and verify successful seeding. Have a TA help you troubleshoot lack of successful seeding.
In FlowSizer, click on Setup under the Home tab and set Maximum Particle Measurement Attempts to 100,000 - this is a temporary change that will buy more time for this demonstration (capture time = MPMA / data rate). Also ensure Screen Update Interval is set to 0 and Time Out is set to 0.
Have one TA stand at the seeder blower control (currently at around 1900rpm), then immediately after hitting Begin Capture have him/her smoothly but quickly ramp the speed down and up a couple of times, before returning to 1900rpm. The Velocity 1 Realtime graph should show this change in speed with a slight delay, until you hit Stop Capture or the 100,000 points have been observed (which they shouldn't have been in this time frame).
Close the seeding valves and set the laser to STANDBY, but otherwise leave everything where it is
Click Save and call the file something meaningful like DemoRun
Click on Data Sets To CSV File and select the relevant metrics, shown in the figure below, before clicking Export
Verify the data exported correctly by navigating in Windows Explorer to the run folder and opening the CSV file. Select the first three columns and click Insert > Scatter to generate a quick preview plot to verify the data makes sense and matches, as shown in the figure below
Return the MPMA setting to 10,000 in FlowSizer Setup
Now you are comfortable with the workflow, it's time to capture real data:
Move the traverse to the zero position (top of the test section)
Open both seeder valves and verify successful seeding. Have a TA help you troubleshoot lack of successful seeding.
Set laser to RUN
Obtain the jet profile data by performing a short capture at each traverse position:
Follow the steps in the previous section to obtain a measurement at the zero position
Having decided on a traverse distance (it doesn't have to be constant throughout the test section... bear in mind the expected profile of the flow in the freestream and the jet), move the traverse to your next position and obtain another measurement.
Repeat Step 2 until you have traversed 155mm downwards (just off the bottom of the test section). You will need to keep a manual list of all traverse positions since this is not logged (or perhaps save your run with a meaningful name).
Go back and repeat the measurements for at least TWO locations (the first point being in the free stream and second in the shear layer), but with different Bandpass Filter/Downmix Frequency settings
Return the filter settings to their proper values, and change the Burst Threshold value to 150 mV and repeat the measurements in the shear layer (to see if the original threshold value was a reasonable choice)
Return the Burst Threshold to its original value, and repeat the shear layer measurement two more times. The first time, turn off (or nearly off) the free stream seeder. The second time turn the free stream seeder back on and turn off (or nearly off) the jet seeder.
Rotate the transceiver by 90 degrees counter-clockwise and acquire vertical velocity component data at the same location used above in the jet shear layer (and additional locations if you wish). Note, you will need to adjust the Bandpass Filter and Downmix Frequency settings to account for the fact that: 1) the vertical velocities are less than the axial velocities and 2) the vertical velocity can be both upward and downward. When you are done taking data, rotate the transceiver back to its original orientation.
Set the laser to STANDBY, close the laser aperture, and close the seed valves, but keep everything else running while you check your data (in case you need to go back and get anything else)
If you didn't do it as you went along, export all your runs to CSV and verify the data looks reasonable. If it doesn't go back and collect any data points.
Sanitize your safety glasses and then assist the TAs in their shutdown procedure if required
LDV system settings (burst threshold, band pass filter, downmix frequency, and transceiver orientation), and vertical measurement position for each run.
Velocity measurements for each run.
Show your calculations for fringe spacing in the laboratory report.
Show your calculations for the filter settings, i.e., what velocities correspond to your chosen filter settings
Table containing the LDV settings and vertical positions for each run.
A plot of the normalized mean axial velocity profile in the jet, versus location, where is the value of in the free stream of the wind tunnel.
A plot of the normalized turbulence intensity profiles in the jet, normalized two ways: first, and second, .
A plot of the instantaneous axial velocities at two locations: in the wind tunnel free stream and in the jet shear layer.
Histograms of the velocity values recorded at interesting points in the flow (note, the histograms produced by the TSI software may not be helpful – you will probably have to make your own histograms).
A comparison between the points taken with different filter, burst threshold and seeder settings.
A comparison of the axial and vertical velocities (mean and rms) acquired at the same jet locations.
This lab covers the use of taps, probes and transducers for measurement of static and stagnation pressures in nominally steady, compressible (subsonic and supersonic) flows. In addition, it includes the use of the schlieren technique to visualize shock and expansion waves in supersonic flows. The various measurements are acquired in a supersonic, blowdown wind tunnel and a supersonic (converging/diverging) nozzle. The experiments also allows the student to experimentally investigate many interesting and important aspects of the behavior of supersonic flows.
Steady pressures are most conveniently and accurately measured using static taps or probes, and stagnation pressure probes. This is the approach used in this laboratory. Static taps and probes, as well as total or stagnation pressure probes (see Figure 1), have been introduced in earlier labs. Typically taps and probes are located quite a distance from the actual pressure transducer, thus they have a relatively slow response. Rapidly fluctuating pressures must be determined using other devices, such as microphones or piezoelectric transducers as covered in previous labs.
Figure 1. Schematic of a) static and b) Pitot pressure probes.
Transducers that can be used to measure pressure include gravitational transducers, such as mercury or oil manometers. Manometers can be of various shapes, including standard U-shaped tubes and straight long vertical tubes attached to a liquid filled reservoir. Manometers measure relative pressures, specifically the pressure difference between the two ends of the tube. For measurement of stagnation pressures in compressible flows, which can easily exceed 1 atm, long tubes are required (even for a dense liquid such as mercury, measurement of Δp=1 atm requires at least a 30 inch long tube).
Figure 2. Schematic of vertical type manometers: a) p < pat , b) p > pat, and c) a manometer bank (with pat=ambient pressure).
For vertical tubes connected to a reservoir, pressures below atmospheric are measured by connecting the tubing from the probe to the upper, open end of the manometer tube while the reservoir is open to the atmosphere (Figure 2a). Thus, the lower the value of the static pressure, the higher the mercury column in the manometer. On the other hand, if the pressure to be measured exceeds atmospheric (e.g., the total pressure in a blowdown type supersonic tunnel), the tubing from the probe is attached to the reservoir and the upper end of the manometer is left open (Figure 2b). If one is dealing with a manometer bank containing a number of tubes, the level in the reservoir drops as the mercury rises in the tubes. This must be compensated for by measuring the drop in level in a reference tube which is open to atmosphere (Figure 2c).
Another family of pressure measurement devices can be categorized as elastic transducers. In these devices, a deflection or deformation accompanying a balance of pressure and elastic forces is used to measure pressure. A classic example of an elastic pressure transducer is the Bourdon tubes (Figure 3) commonly found in pointer type pressure gages. In these devices an oval section tube is initially coiled into a circular arc. As a pressure is applied to the tube, the oval section tends to become more circular in cross-section. Since the inner and outer lengths of the tubes remain approximately the same as their initial values, the primary result of the applied pressure is an uncoiling of the arc. The uncoiling is coupled to the motion of a pointer, which is used to determine the pressure.
Figure 3. Basic Bourdon-type pressure transducer.
Another type of elastic transducer is the diaphragm type, where the pressure force causes a diaphragm to deflect. Examples of this type include Barocell/Barotron type transducers, in which the pressure deflects a thin diaphragm that forms one plate of a capacitor. Different pressures yield different capacitances that are converted to electrical voltages (and read using a voltmeter in this lab). Barocells and Barotrons typically provide very sensitive measurements of pressure, e.g., Torr/volt, with the calibration (sensitivity) usually supplied by the manufacturer. Other methods can be used to measure the change in the diaphragm, for example resistance strain gages can be applied to the diaphragm surface, and the measured strain can be related to the applied pressure through calibration. Relatively inexpensive transducers can be made by using semiconductor materials. In this case, the semiconductor resistors are “written” as a bridge circuit directly onto a substrate (e.g., silicon) that acts as the diaphragm. The strain on the semiconductor results in a change in semiconductor resistance; this is known as the piezoresistive effect. The change in semiconductor resistance is analogous to the change in metal resistors (recall the strain gauges used in the force balance experiments), except in the latter, the change in resistance is primarily due to the change in the metal resistor’s cross-sectional area as it is strained. For semiconductor materials, the resistance change is related to other changes in the internal structure of the semiconductor. This type of silicon diaphragm transducer will be used to make differential pressure measurements in the converging-diverging nozzle experiment.
Sudden changes in the density of a gas and the resulting gradient in refractive index can be visualized using the schlieren technique. Such steep refractive index gradients exist, for example, in flames or in shock waves. In this laboratory, shock waves, whose properties will be discussed in the next section, are visualized using schlieren imaging. In this technique (see Figure 4), the light from a source is allowed to expand and is then collimated using a large diameter, long focal length (typically biconvex) lens.
Figure 4. Schematic of a typical schlieren setup.
The resulting parallel beam is then passed through the test section of the wind tunnel before being refocused using a second biconvex lens of similar optical properties to those of the first. A stop placed at the focal point of the second lens blocks all of the undeflected light. However, any light rays that have been refracted in the test section by a refractive index gradient caused by, for example, a shock wave are no longer parallel to the optical axis of the system. These rays will, therefore, be focused at a different location in the focal plane of the second lens and will, thus, bypass the schlieren stop. A focusing lens is then used to create an image of the test region on a screen or photographic plate using only the refracted light. This image displays only those regions of the flow in which a steep refractive index gradient exists, i.e., the shock wave.
A brief description of compressible flows (summarizing important details covered in AE 2010) is given below. It is important to note that inviscid flow will be assumed throughout this discussion. When a gas, such as room temperature air, flows at velocities greater than approximately 100 m/s (or for Mach number M > 0.3), the density changes of the gas due to changing Mach number become significant, i.e., the flow becomes compressible. This results in some remarkable phenomena, especially at supersonic speeds. For example for subsonic flow in a nozzle, M increases as the cross-sectional area of the nozzle decreases; for supersonic flow this trend reverses. Furthermore, it turns out that an adiabatic flow can only increase from subsonic to supersonic speeds if the transition from M<1 to M>1, (i.e., M=1) occurs at the throat (minimum area) of a nozzle.
Bernoulli's equation, which was derived assuming constant density is no longer valid for compressible flow. Instead, the relationship between the static and stagnation pressures, e.g., , as a function of Mach number may be obtained for a calorically perfect gas (constant specific heats) from the following (more general) expression.
(1)
where is the specific heat ratio. As long as the flow remains isentropic, the stagnation pressure is constant everywhere. The static pressure, on the other hand, decreases as the Mach number increases since more potential energy, which gives rise to the static pressure, is now converted to kinetic energy. This is true for subsonic as well as supersonic flow.
Figure 5. Schematic of the converging-diverging nozzle showing location of pressure probes and taps (throat at ).
Let us now consider the flow in a converging-diverging nozzle (see Figure 5) connected to a high pressure supply at one end and open to the atmosphere at the other. As the pressure in the supply (reservoir pressure) is increased, air begins to flow through the nozzle. As long as the flow in the nozzle is everywhere subsonic, the Mach number increases in the converging part (region A) and decreases in the diverging part (region B) of the nozzle. The static pressure level in the nozzle behaves inversely as the magnitude of the Mach number. Thus for subsonic flow in the nozzle, the static pressure drops in the converging section and rises in the diverging section. At the exit of the nozzle, the flow should reach equilibrium with the back pressure (given by the local atmospheric pressure in this example). As the reservoir pressure is increased further, the mass flow rate through the nozzle increases, causing a change in Mach number and, thus, in static pressure. The Mach number in the converging part of the nozzle can keep increasing in this manner until the Mach number at the nozzle throat reaches unity. The nozzle is then called “choked”. The Mach number in region A can no longer be affected by increasing the reservoir pressure (or decreasing the back pressure), and the mass flow rate, for fixed upstream stagnation conditions, is a maximum once the flow is choked.
Increasing the reservoir to back pressure ratio above the value that corresponds to choking does, however, affect the flow conditions in region B. Beyond the choking condition, the flow downstream of the throat begins to go supersonic and increases in Mach number as the nozzle flow expands. However, is not yet high enough to result in completely supersonic (isentropic) flow throughout the entire region B. The flow adjusts to these conditions by suddenly reverting back to subsonic flow in a normal shock somewhere in the expanding part of the nozzle. Across this shock wave the static pressure rises. Behind the shock the flow is subsonic, and, therefore, decreases in Mach number as the nozzle continues to increase in area. The static pressure rises correspondingly until the back pressure is reached at the exit of the nozzle. As is raised still further, the position of the shock wave moves towards the nozzle exit, until eventually there is no shock in the nozzle.
The presence of the shock wave changes all flow conditions across it except the stagnation temperature (or, more precisely, the stagnation enthalpy) since across the shock the flow is adiabatic. The extent of the influence of the shock upon the flow conditions depends upon the Mach number of the flow going into the shock. For thermally and calorically perfect gases, the ratio of the static pressures across the shockas a function of Mach number ahead of the shockare given by
(2)
and the stagnation pressure ratio is
(3)
where we have assumed a stationary shock for the stagnation pressure ratio.
Figure 6. Schematic of shock on: a) probe and b) wedge in supersonic flow (θ = shock half angle, δ = wedge half angle).
If a supersonic flow encounters a solid body, part of the flow must be decelerated to stagnation conditions. Since this necessarily involves a transition from supersonic to subsonic flow, a shock stands ahead of the body. This shock provides the mechanism for the transition. In the case of a thin probe in supersonic flow a small, normal shock stands ahead of the probe tip (Figure 6a). A conically shaped bow shock trails from the normal shock. The bow shock weakens as one moves away from its leading edge and, eventually, turns into a Mach wave across which the flow properties no longer change significantly. Therefore, the stagnation pressure measured using a Pitot probe in supersonic flow is that behind a normal shock. In a static probe, on the other hand, the orifice is located far downstream of the probe tip. The effect of the normal shock is then no longer felt and the static pressure measured is essentially equal to that ahead of the shock. If the probe is replaced by a wedge (Figure 6b) a pair of oblique shocks is formed that attach to the tip of the wedge. The relationship between the flow Mach number and the half angles of the shock and of the wedge are plotted in Figure 7.*
*Similar plots can be found in most texts on compressible flow.
Figure 7. Variation of shock-wave angle with flow-deflection angle for various upstream Mach numbers for a thermal and calorically perfect gas with γ = 1.4.
This set of experiments is to be carried out in a converging-diverging (CD) nozzle made of Plexiglas (see Figure 5). The upstream side of the nozzle is attached to a high pressure supply (regulated so as not to exceed ~50 psig); the downstream side is connected to a settling chamber and an acoustic muffler before exiting into the room. Between the nozzle and the high pressure supply is a manual valve. The stagnation pressure entering the nozzle is controlled by the setting of this valve (and the pressure downstream of the valve can be monitored using a simple Bourdon tube type gauge). The nozzle is instrumented with two Pitot probes along the axis and eight static pressure taps along the wall. The first Pitot probe is located at the nozzle throat. The second probe is located near the last static pressure port in the nozzle (station 7).
Figure 5. Schematic of the converging-diverging nozzle showing location of pressure probes and taps (throat at).
As shown in Figure 5, one static pressure tap is located ahead of the nozzle throat, the second at the throat and five tops are located at one half inch intervals along the expanding part of the nozzle. Tap number eight measures the back pressure. The ten pressures will be measured by ten piezoresistive, silicon diaphragm transducers and recorded by a computer data acquisition system. Since some of the pressure gauges can be damaged by operation outside their pressure range, please carefully listed to any guidelines given by the TA’s with respect to the maximum pressure you should allow the CD inlet to reach.
In this facility, air from a large tank at approximately 120 psi is passed through a pressure regulator and a large, manual (butterfly) valve and into a supersonic tunnel of rectangular cross-section. The tunnel (Figure 8) consists, essentially, of a nozzle (2" × 1.778" at its throat) and a test section (2" × 3").
Figure 8. Schematic of supersonic blowdown tunnel.
The test section walls are fitted with two parallel windows in order to provide optical access. The tunnel is instrumented with a Pitot probe upstream of the nozzle and a removable Pitot-static probe that mounts in the test section. The orifice of the static pressure probe is aligned with the tip of the Pitot probe in the test section. The Pitot probe upstream of the nozzle is connected to a pressure gauge (which records gage pressure). The downstream Pitot probe can be connected to the Baratron for comparison to the ambient pressure. Similarly, the static pressure probe can be connected to the Baratron while the other side is open to atmosphere. Atmospheric pressure is measured using an electronic barometer. The probes in the test section may be removed and replaced by a wedge or by a solid body of some other configuration.
Figure 9. Schematic of schlieren setup attached to “2-d” blowdown windtunnel. The graduated color filter has a dark red pattern in the center that acts as the schlieren stop.
A schlieren system (Figure 9) is configured around the test section of the tunnel. A single lens is used to perform both the focusing of the collimated light and to image the test section on the screen. Also, a bicolor filter with an opaque band in the middle is used to produce the schlieren stop. Rays that pass above and below the focal point of the second lens pass through different colored filters. Thus rays that were deflected upwards will show up in one color in the image, while rays that were deflected downward will be a different color. Although the tunnel is fed from a large tank, running times are limited. It is, therefore, important to coordinate the measurements to be made before starting the tunnel. It is a good idea to watch the upstream stagnation pressure to make sure that it is not dropping with time (which would happen if you were overdraining the supply tanks). Since the Mach number in the tunnel depends only upon the area ratio of nozzle to test section (as long as the upstream pressure is sufficient to choke the nozzle), the test section Mach number is not affected by any pressure drop, but your pressure measurements will change.
The compressed air supply for this experiment is not infinite and the flow rate through the tunnels is substantial. As such, keep your runs to a minimum else you run the risk of running the tank down faster than the compressor can keep up with. If this occurs, you will need to wait for the compressor to charge the tank before being able to complete your experiments. To gauge the capacity of the tank open the valve to the blowdown tunnel and check the first pressure gauge in the line; it will read ~120 psi when full, and the compressor will kick in at ~85 psi. Pressures much below this may result in inability to complete the experiment.
The pneumatic fittings used in this experiment are all "push-to-connect" style. To make a connection firmly insert the tube until you feel the inner sleeve give and make the connection. To disconnect, do not pull on the tube without first pushing down the black plastic ring around the tube.
Take a tour of the equipment with the TAs. In particular you will look discuss:
The large grey tank outside the door, which feeds our experiments. This tank is charged by a compressor (that you won't see), automatically kicking in when the pressure drops to ~80 PSI.
The pneumatic system feeding both of our experiments from the wall (and ultimately the tank).
The CD Nozzle experiment, making sure to look at the nozzle directly and comparing it to the diagram.
The Blowdown experiment and its removable inserts.
PPE and other safety items.
The LabView VI.
Preparing the compressed air system
Close the small black-handled calibration valve labeled "AIR CAL" and the large red-handled valve labeled "AIR EXP CD".
Open the three shut-off valves on the wall, labeled "AIR SOV 1", "AIR SOV 2", and "AIR SOV 3". This opens the main compressed air feed from the large grey tank outside the door next to our experiment, and energizes both the CD nozzle experiment and blowdown tunnel for later.
Verify, using the pressure gauge labeled "PRESS CD", that the CD nozzle experiment is energized (around 80psi).
Check, using the pressure gauge on the blowdown tunnel labeled "PRESSURE BDN IN", what the main compressed air storage tank pressure is.
If the tank is fully charged it should read around 120psi
As the experiments are run this reading will gradually drop
The tank compressor will not kick in until this reads around 80psi, at which point it will take 15-20 minutes to charge the tank back up to 120psi
Keep an eye on the compressed air supply pressure (PRESS BDN IN) throughout the course of this lab... if you start losing velocity in the tunnels it's likely because the tank pressure is getting too low and you will need to wait for it to charge. Keep your runs as short as possible... wasted time is wasted air and then waiting time!
Preparing the instrumentation
Turn on the big screen and set the it so that: Scale = 200%, Display Resolution = 3840 x 2160, Multiple Displays = Extend These Displays.
Navigate to D:\AE3610 Fall 2021\Supersonic Flow
and open the Supersonic
VI in LabView. Drag it over to the big screen.
Power on the transducer box and check that the VI registers a voltage change
Calibrating the transducers
Connect all the tubes coming from the transducers to the calibration box. Check the calibration box labels for what goes where. Do not remove the black compression fittings from the calibration box tubes, or remove the tubes from the transducers.
!!! Put on your eye protection !!! (ear protection not required yet)
Follow the instructions in the VI for performing the calibration
Take a screenshot/picture of the final NULL and SCALE values for all 10 transducers, or manually record them, for later
Gather some "calibration check" data to verify your VI calibration values against the manometer as follows: By clicking "TAKE DATA" on the VI, record data using the VI at 3 different manometer pressures that you should also make a manual note of. As you go through this process, sanity check that the VI and manometer match. By slowly varying the black-handled calibration valve, the pressures you should evaluate are:
Whatever pressure you calibrated at (it should still be there)
Something roughly half of the previous pressure (it's not critical what this pressure is so don't waste time dialing it in)
0 psi (AIR CAL valve fully closed)
Open the save file to ensure your data has saved correctly. Do this in Notepad++ (to ensure there is no write conflict with the VI)
Make a manual note of the atmospheric pressure using the weather station on the desk.
Performing the experiment
Make sure AIR CAL is fully closed.
Connect the transducers to the correct pressure taps in the experiment (P1 to P7 followed by Pb go in order from left to right when viewed from the side of the test section nearest the gas turbine). Take care to not overly stress the tubes coming out of the CD nozzle.
!!! Put on your ear and eye protection !!! You are about to run the experiment, during which substantial air will flow, so be sure to read the following instructions carefully and discuss your plan of action before proceeding from here. You are welcome to, and indeed should, stop the tunnel at any time by closing the AIR EXP CD, if you wish to discuss something you've observed with your group, or want to plan on re-taking data. Remember: wasted time with the air running is wasted air!
Get a feel, without logging any data, for how sensitive the valve is and how pressures in the nozzle will change throughout the whole range of operation.
Do this by slowly opening AIR EXP CD and observing the VI pressures in real-time.
During this step discuss what the flow speed regime is, where the shock wave is, where the flow is choked, etc.
Identify what positions you are required to take data in the VI data
See if you can see the shock wave by eye (not always possible)
Now we will take data at the required points. Slowly open the AIR EXP CD valve to allow subsonic flow through the CD nozzle. Use the VI to record pressures in the nozzle at a high enough speed to see the subsonic pressure profile clearly, but without going supersonic
Open the valve further until the flow is just choked, i.e., throat has just reached Mach 1. You may also hear a distinct change in the noise from the nozzle. Use the VI to record the pressures.
Continue opening the valve until a shock is observed between stations 4 and 5. You will be able to detect the shock location by a change in sign in the pressure gradient in region B of the nozzle (refer to the figure in the VI for the regions). Once again, record all pressures with the VI.
Continue opening the valve until the shock passes station 6. The flow in most of the expanding part of the nozzle should now be supersonic. Once again, record all pressures using the VI. Note that, just before and during this phase, you will see P1 and P2 peak at ~20 psi and no longer rise. These transducers have a maximum safe pressure rating of 50 psi, but can only read valid pressures up to 20 psi, so they will seemingly stop reading at this point... a phenomenon called "saturation". Whilst saturated we will not know what the real pressure is so be sure to not exceed the valve position that creates a shock just past station 6.
Repeat steps 5.5 - 5.8. This will provide data to assess repeatability.
Close AIR EXP CD.
Check your data
Click Done on the VI to save all data.
Open the file you selected as your save file in MS Excel. If you didn't give it a file extension when you selected the save file you will need to add it now by right-clicking on the file, clicking Rename, and then adding ".xls" to the end of the filename.
Delete any rows you accidentally captured data at, or decided to over-write.
With the help of the TAs, visually observe that all data was saved properly and is of the rough order of magnitude of what it is supposed to be.
If you are missing any data or it is not correct, go back and collect it now.
Repeat Step 6.1-6.5 until you are satisfied the data is acceptable.
Shutting down the experiment
Purge the compressed air (we don't want high pressure sat in a flexible hose in case it comes loose and disconnects, potentially causing harm to someone)
Close AIR SOV 1
Open AIR EXP CD slowly, until there is steady flow to purge the line
Once the air stops flowing and CD PRESSURE reads 0, close AIR SOV 2 and AIR EXP CD
Power down the transducer box
Make a plot of the 3 pressure readings from the manometer ("true") versus the VI readings obtained using the calibration NULL and SCALE readings ("measured"). Discuss the accuracy the calibration and how it could be improved.
Convert all of the transducer readings to absolute pressure. These will be required for later calculations and plots.
Calculate the Mach numbers at stations 1 and 2 from the static and stagnation pressures determined at these locations for the following three conditions:
subsonic flow throughout the nozzle
the “just choked” condition
the condition where the shock stands between stations 4 and 5.
Calculate the Mach number at station 6 for supersonic flow in two ways:
using the static pressure at station 6 and a suitable stagnation pressure; and
using the stagnation pressures as measured by the Pitot probes located at the throat (station 2) and at station 7.
Note - be sure to use the correct value of the stagnation pressure in these calculations. It will be helpful to sketch for yourself a diagram of the nozzle indicating the position of the pressure taps and probes and the location of the shock wave.
Make a table of the sensitivity (scale) and zero offset values for each pressure transducer.
Make a table listing:
atmospheric pressure
the stagnation pressure measured using the two Pitot probes for the cases of:
subsonic flow
a shock between stations 4 and 5
supersonic flow through at least station 6.
Make a table listing the Mach numbers at stations 1 and 2 for the three conditions specified under Data Reduction step 3
Make a table listing the Mach number at station 6 as calculated by the two methods given under Data Reduction 4
Make a graph with axial distance as the abscissa and the ratio of local static pressure to supply stagnation pressure as the ordinate. On this single graph, plot the results for the static pressure measurements for the conditions of:
subsonic flow
a shock between stations 4 and 5
supersonic flow through at least station 6
Using different symbols, plot the repeated data on this graph as well. This will show the quality of the repeatability.
Make a plot like that described in step 5 above, except plot Mach number as the ordinate.
The compressed air supply for this experiment is not infinite and the flow rate through the tunnels is substantial. As such, keep your runs to a minimum else you run the risk of running the tank down faster than the compressor can keep up with. If this occurs, you will need to wait for the compressor to charge the tank before being able to complete your experiments. To gauge the capacity of the tank open the valve to the blowdown tunnel and check the first pressure gauge in the line; it will read ~120 psi when full, and the compressor will kick in at ~85 psi. Pressures much below this may result in inability to complete the experiment.
The pneumatic fittings used in this experiment are all "push-to-connect" style. To make a connection firmly insert the tube until you feel the inner sleeve give and make the connection. To disconnect, do not pull on the tube without first pushing down the black plastic ring around the tube.
Preparing the compressed air system
Ensure the main blowdown tunnel control valve, AIR EXP BDN, is fully closed
Ensure AIR SOV 3 is still open from when you opened it earlier
Repeat CD Nozzle Procedure Step 2.4 to verify air tank status
Ensure AIR REG 1 and AIR REG 2 are OPEN. And that AIR REG 3 is CLOSED.
Pitot tube experiments
Install the pitot tube plate into the test section
Connect the high pressure port of the manometer to the pitot tube stagnation port (the manometer low pressure port should be open to the atmosphere). Turn on manometer, zero, and change units to psi.
!!! Read the following instructions fully before commencing - wasted time is wasted air !!!
Ensure you are wearing your eye and ear PPE. Gradually open the main blowdown tunnel control valve until the upstream pressure, PRESS BDN US, reads around 14 psi. Make a note of the exact pressure obtained.
Record the pitot stagnation pressure on the manometer.
Working quickly, but not rushing, switch the manometer tube to the static pressure port and record the pressure once it's settled
Fully close AIR EXP BDN
Make a note of the atmospheric pressure using the weather station on the desk
Wedge experiments
Remove the pitot tube plate and replace it with the wedge plate
Connect the P1 port of the manometer to the TOP static port (the manometer P2 port should be open to the atmosphere).
Enable cooling air to the light source by opening the small black-handled valve
Have a TA turn on the Schlieren light system and wait for the lamp to warm up (avoid switching the light source on and off unnecessarily since this will shorten its life and may extend your time in the lab if it doesn't re-illuminate quickly)
Using your camera phone take a picture of the wedge's shadow on the screen. Be sure to get orthogonal to the screen and fill the frame as much as possible. Take multiple shots then select the best.
!!! Read the following instructions fully before commencing - wasted time is wasted air !!!
Position yourself around the screen to observe the wedge's shadow
Ensure you are wearing your eye and ear PPE
Have a TA slowly open the main control valve whilst you watch the screen; at some point you will see a new shadow appear around the wedge; the flow is now supersonic and this new shadow is the Mach cone. Tell the TA to stop opening the valve.
Record the following things:
Upstream pressure from the PRESS BDN US gauge near the TA
Static pressure on the manometer
A new picture of the Schlieren screen with the Mach cone visible. Be sure to get orthogonal to the screen and fill the frame as much as possible. Take multiple shots then select the best.
Close AIR EXP BDN.
Shutting down the experiment
Have the TA turn off the Schlieren lamp and wait for it to fully stop shining, at which point turn off the cooling air.
Close AIR SOV 3 valve.
Open AIR EXP BDN on the blowdown tunnel and wait for the air to fully purge through the wind tunnel (we don't want to leave any portion of the experiment charged with high pressure air!)
Close AIR EXP BDN and the Schlieren lamp cooling valve
Disconnect the manometer and pack it up in its case
Checking data before you leave
Have the TA sanity check your blowdown values before you leave
If you have time, draw your blowdown tubing schematic now as a group so the TAs can sign off on it. If you don't have time, take pictures and ask the TAs questions as required
Calculate the test section Mach number from the ratio of static to stagnation pressure in the test section. Be very careful to use the correct, measured stagnation pressure. A rough sketch of the position of the probes and the location of the shock may help you.
Calculate the test section Mach number using the measured stagnation pressures and the shock equations (listed in the Background section) or appropriate shock tables (preferred method).
Determine the Mach number in the test section using the values of the half angles of the wedge and the shock that you measured. Use Figure 7 or similar graphs from your compressible flow texts.
Make a schematic of the various configurations of tubing used in the pressure measurements.
Make a table listing the test section Mach number as calculated by using:
the ratio of static to stagnation pressure in the test section
the measured stagnation pressures in the test sections
the shock angle on the wedge
This page details information about the AE 3610 Final Project
The purpose of this project is to introduce you to experimental design by allowing you to design and execute your own experiment. You will be given the unique opportunity of designing your own experiment within the realm of fluid and solid mechanics from beginning-to-end. In AE 2610 and AE 3610 thus far, you have performed prescribed experiments where the problem definition, experimental design, experimental procedure, and data reduction process was taken care of for you. In this project, you will perform all of these aspects of the experimental process yourselves. These aspects include but are not limited to:
Conceptualization
Experimental goals and motivation
Experimental procedure
Fabrication of test articles
Testing
Obtaining raw data and processing data
Reporting and presentation of findings
This project will span almost the entire semester. You will be working in assigned groups of 3-5 people and will be assigned either the topic of solid mechanics or fluid mechanics. The ultimate goals of this project are to expose you to self-guided experimental design, help you to develop your engineering curiosity by molding the way you approach engineering problems, implement the theoretical knowledge you have gained thus far in your AE courses, implement the experimental knowledge you gained from AE 2610 and 3610, and introduce you to fabrication and working with hardware.
As mentioned before, this project will be as self-guided as possible to ensure you guys are thinking through the experimental process on your own. You will decide what to test, figure out a way to do it on your own, and execute an experiment to get data and present findings. You are expected to try, fail, persist, and if necessary, pivot. You will have the course professors and TAs to guide you through this process, but we will encourage you to try to do what you can on your own!
Once you have picked a topic, your group will submit a project proposal containing the information in the “” section below. Upon submission of this proposal, the feasibility of your project will be assessed by a technical team consisting of the course professor, the lab manager, and the TA assigned to your section. This technical team will give you feedback and give you advice to ensure that your project is feasible and has an appropriate level of complexity while maintaining the overarching goal of the project. Once your project has been approved by the technical review team, you will begin working on your project.
There are three main components you are required to consider when designing your experiments:
Creativity and complexity of the experiment – Don’t redo a lab you’ve already done in 2610/3610
Design and fabrication of your test specimens and other fixtures
Testing process and data reduction
When designing your experiments, think about the trade-offs between each of these three components. You may realize that your project is a bit simple but is very fabrication heavy. You may realize that your project may not be too complex and fabrication is simple, but your data reduction process will be extensive. These are all acceptable! Experiments are different and focusing on one part versus the other is acceptable. What is NOT acceptable is skipping one of these three required components altogether.
Please know that this project is to be within the scope of this course—we are not expecting you to be doing experimentation and documentation worthy of journal articles (although if you do, we obviously won’t stop you)! From a course perspective, we want you to gain good exposure to designing experiments, working with hardware, and understanding and developing the experimental process through your own expertise, though your TAs and course instructor will be available to help you.
Furthermore, know your project may not work. This is okay—experimentation does not always play nice. We are interested in your thought process and effort more than the results. If your project fails, you are expected to pivot if necessary, and put effort into figuring out why your experiment didn’t work. Despite your experiment failing, you can still earn a great grade if you put the thought and effort into your project! The graders are aware of this.
In order for us to fully assess the feasibility and support needs for your experiment, your proposal needs to be accurately and thoroughly communicated. As such, ensure it includes the following contents at a minimum. An example project proposal and a proposal template are provided at the end of this section to aid you in creating your proposals. Include:
Your section number, fluids/structures, and list of section members, semester and year
A short summary of what the over-arching goal of your experiment is (1-2 sentences)
An extended section of prose that contextualizes and motivates the experiment have chosen
What your dependent variables are and how your plan to measure each dependent variable
What your independent variables are and how you plan to control each independent variable
An itemized list of specific equipment that you will use, and who will provide/fabricate those items. List all:
Lab spaces and associated lab equipment/apparatuses
Instrumentation
Sensors
Tools
Test articles and fixtures
What software you will use to execute the experiment or acquire data
How you plan to mount your test articles, and any modifications that are required to facilitate that mounting
A Team Responsible Itemized Bill Of Materials (BOM) for every part and consumable that needs to be purchased, including information on:
Item Descriptions
Quantities of each item
Unit costs of each item
Extended costs (unit cost * quantity) of each item
The vendor name with a hyperlink to the specific URL for each item
The Total Cost of your order
Remember your budget is $100
We do not pay sales tax!
Note: It is best to do this in table format
Item Descriptions
The amount of each consumable you will use
Who will be providing the consumable item
Note: It is best to do this in table format
What fabrication processes you will need to perform, where those processes will happen, how those processes will happen, and who will perform them
Note: you MUST be machine shop trained to use the AE Machine Shop!
What raw data you will be collecting and how it will be post-processed and reduced
A high-level timeline outlining fabrication, testing, data processing, report writing, etc.
A preliminary experimental procedure that outlines how exactly you will execute your test
You are free to format the above information according to your judgement, but it is not necessary to go over the top or use a lot of prose. Only the bare minimum to communicate the above in a readable fashion is required; bullet points and tables should cover most of it.
Below is an example proposal in pdf format:
Below is an example proposal in docx format. You may use this as a template to create your own.
Each group will be provided with a discretionary $100 fund allocation that can be spent at either McMaster-Carr or Home Depot (Midtown branch ONLY). This will come in the form of a one-time purchase in the days after the final project plan and BOM list is finalized. All items on the list will be picked up in person by GTAE staff once and once only; it is on you to make sure your items are verified as in stock at the aforementioned locations. Items will be made available for pickup at the same time for all sections. An announcement will be made when items are available for pick-up. Any funds not used during this one-time purchase will be forever lost, so use it wisely. You will not be charged for sales tax. Purchases from other vendors or after this one-time purchase are permitted, but no funds or purchasing support for these items will be provided--you must purchase these yourselves.
In addition to discretionary funds for each section, GTAE also provides certain raw materials via the Aero Maker Space. The exact list of raw materials should be checked each semester with the AMS directly, but typically includes 3D printer filament, various bits of plastic and wood, and lots of consumables.
Ensure your shopping cart is full of all the items you wish to purchase
Go to your email account and the URL for your shopping cart will be found at the very bottom of the email sent to you by McMaster-Carr
Create a hyperlink in your proposal document called "McMaster-Carr Shopping Cart Link", pasting the URL in from above (don't just paste the giant URL in directly)
Click on the hyperlink to ensure it works properly
Ensure your shopping cart is full of all the items you wish to purchase
Create a hyperlink in your proposal document called "Home Depot Shopping Cart Link", pasting the URL in from above (don't just paste the giant URL in directly)
Click on the hyperlink to ensure it works properly
Prior to proposal sign-off, each group's initial proposal will be reviewed by a technical review team including the Lab Manager, an assigned TA who will remain with you throughout the rest of the semester, and the professors. Thereafter, your section will meet with the technical review team during your assigned lab time to have a dedicated consultancy regarding your plan and how to execute it. After final signoff, teams will have continued access to their assigned TAs for assistance during assigned lab times and, to a limited degree, other times throughout the week.
Finally, if your test plan requires something that will be useful to future semesters (such as a particular turntable mount, a tool, or a load frame fixture), GT-AE *may* provide engineering and financial support for these. Any such item should be proposed to the technical review team, but there is no implied guarantee of such support.
In all cases, please remember that all resources outside of your section are subject to real-life constraints and are not solely dedicated to your project. As such, there is no implied guarantee that any of these resources are available to you when and how you need them. Where possible, we recommend that you aim to be as self-sufficient as possible, to maximize your chance of success.
These lists define the available equipment provided by the labs for your use. Some other equipment may be available but not listed, so if you have a unique need please include this in your project proposal and we *may* be able to accommodate it.
Low Turbulence Wind Tunnel, including:
The Baratron pressure transducer
The FlowKinetics flow meter
The 6" pitot traverse, mounted in one of two locations:
Location 1: traversing vertically from the floor 12" away from the near-side tunnel wall
Location 2: traversing horizontally 7" above the wind tunnel floor
The ceiling pitot-static probe
This tunnel reaches a max speed of 35 mph
Test section area is approximately 42" x 42" including the boundary layer
Pitot-static probe
24-column liquid manometer
Wake rake
Boundary layer mouse
3-axis sting balance
This tunnel reaches a max speed of 120 mph
Test section area is approximately 12" x 12" including the boundary layer
Other equipment:
Handheld hot wire probe
Thermal imaging camera
A small toolbox with miscellaneous tools including those for marking and measuring
Calibrated masses
A variety of clamps
Instron 5982 load frame, including:
1 x 350 Ohm Quarter-Bridge Strain Gauge Adapter
2 x 120 Ohm Quarter-Bridge Strain Gauge Adapter
Extensometer
100kN load cell
Fixtures from your AE2610/AE3610 experiments:
Buckling V-Grooves (1" groove length)
3-point and 4-point bending (12" max specimen length)
Dogbone clamps (2" x 2" clamping area and 5/16" max opening width)
Vishay 7000 strain gauge data acquisition system + Vishay software
Aramis 3D DIC system, configured for 4x4" viewing window (the area of interest of your test article must fit inside this window)
Other equipment:
Thermal imaging camera
A small toolbox with miscellaneous small tools including those for marking and measuring
Calibrated masses
A variety of clamps
An assortment of tapes:
Gorilla Tape
Duct Tape
Scotch Tape
Clear Packaging Tape
Aluminum Foil HVAC Tape
Masking Tape
Blue Painters Tape
Electrical Tape
An assortment of adhesives:
JB Weld
JB KwikWeld
Super Glue
Wood Glue
All-Purpose Glue
Ramboard and Packing Paper
Fluids Specific Consumables:
Tubing for the pressure transducers
Propylene glycol for the smoke generator
Structures Specific Consumables:
Uniaxial strain gauges (up to 4 per group)
Strain gauge application chemicals and equipment
DIC speckle application:
Self-adhesive printer labels
Label printer + speckle generation software
Matte/flat white spray paint for priming the specimen
Black spray paint and ink for the speckles
Toothbrushes for application implementation
A universal wing mount was created for fluids groups to use in the Low Turbulence Wind Tunnel. This mount will act as both a structural spar for wings to minimize flexure and torsion, and as an adapter to fix the wing to the load cell adapter plate. It is made from 1/2" thick Aluminum 6061.
The wing mount can be mounted onto the load cell adapter plate by aligning the dowels on the load cell adapter plate to the dowel holes on the wing mount. The wing mount will then be secured to the load cell adapter plate using 5/16" shoulder bolts, similar to how the AE 2610 Subsonic Experiment Wing was mounted.
The wing mount has two threaded holes for 1/2 NPT galvanized iron pipe. Two 1/2 NPT pipes of varying lengths will be threaded into these and will act as a spar support to minimize wing flexure and torsion. Two pipe lengths will be available for use:
9-3/4 inch
25 inch
Please size your wings accordingly. Note that wing spar shall not protrude from the end of your wings (e.g., a 15" spanwise wing cannot utilize the 25" spar and you must use the 9-3/4" spar provided). An isometric view of the wing mount is shown in the figure below featuring the 9-3/4" spar.
Shown below is a top view of the wing mount with the necessary dimensions for your projects. Please use these dimensions when designing the ribbing for your wing, as the holes for the spars will need to be cut into the ribs. Once constructed, your wing should simply slide onto the spars and into place for testing. The centers of the spars are 1.75 inches apart and the spars themselves have an outer diameter of 0.848"+/-0.005" and an inner diameter of 0.625". Note that there is a tolerance for the thickness of the spars. All dimensions are in inches.
There are two methods available for use for this project. One will be used for static load testing and the other will be used for tensile testing. Depending on the scope of your project, you may select either one for use.
This is the procedure to use these methods:
Decide if you will be using the extensometer or strain gauges for your experiment. You will need to use the appropriate transducers. The panel is located on the bottom left of the machine near the rear.
The open panel looks like this
Do not touch/remove "120 OHM A" transducer from the "Strain 1" port
Transducers must be swapped BEFORE turning the Instron machine on
Insert the appropriate transducers depending on whether you are using strain gauges or the extensometer for your experiment
Using 1 OR 2 strain gauges:
Open the panel on the back bottom left of the Instron machine.
Ensure that the "120 OHM B" transducer is plugged into the "Strain 2" port. Swap transducers if necessary.
Using only the extensometer OR 1 extensometer and 1 strain gauge:
Open the panel on the back bottom left of the Instron machine.
Ensure that the Extensometer transducer is plugged into the "Strain 2" port. Swap transducers if necessary.
Power on the Instron machine. The switch to turn it on is on the bottom right of the machine. Allow at least 10 minutes for the machine to warm up.
While the Instron is warming up, power on the computer. The computer used for this machine is the monitor attached to the right of the machine. The power on button is a small circular button located between the USB ports and the power cable on the back of the monitor.
Once the Instron machine is warmed up, open the Bluehill Universal program from the Desktop. The home screen is shown below.
Install your specimen using one of three available fixtures:
V-Groove Fixture
4-Point-Bending Fixture
Dogbone Clamp Fixture
Use the jog up and jog down buttons on the Instron machine control panel move the crosshead up and down. Use the Fine Position dial to make small changes to the vertical position of the crosshead. Take care not to jam the fixtures against each other or accidentally yield your specimen--Remember there is a lag time to releasing the jog up and down buttons.
Ensure that you have installed your test specimen correctly into the Instron machine. Refer to previous experiments if you need a refresher on how to install your specimen.
Connect your strain gauges to the strain gauge adapter modules (120 Ohm A and 120 Ohm B). If you are using only one strain gauge, use only the 120 OHM A adapter.
Be sure to note which strain gauge is connected to which adapter module in the event that you are utilizing both adapters.
Click Test
on the Bluehill home screen. The following screen should open.
The static testing method is a manual load application method. This method can acquire data for up to two strain gauges. It is very similar to the Bucking experiment method used in the Bending and Buckling lab from AE 3610.
The jog up, jog down, and fine position dials must be used to apply a load manually onto your test specimen. The strain data must also be recorded manually, as was the case with the buckling experiment).
From the Test screen, select the red Browse methods
button on the right of the page.
Select the appropriate method
If using extensometer
Select AELabs>AE3610>AE3610 Project>AE3610StaticLoading-Extens.im_tens
If using only strain gauges
Select AELabs>AE3610>AE3610 Project>AE3610StaticLoading-Gauges.im_tens
Wait for the sample to initialize, and then select the red OK
button. The following method screen will open
Hit Balance all
on the bottom of the screen. All four of the live displays on the top of the screen should read zero or close to 0. If the strain starts drifting, ensure that the wires are connected to the adapter correctly and hit Balance all
again.
Perform your test by loading your specimen. Load your specimen using the jog up, jog down, and fine position buttons on the control panel.
Manually record all your data. There is no report exported for this method. Make sure you sanity check your data before leaving.
Remove your specimen from the Instron and power it down by turning off the switch on the bottom right of the machine.
Turn off the computer by pressing the small circular button between the power cable and the USB ports on the bottom of the monitor.
The tensile test method is a load application method wherein load an increasing amount of load is applied to the specimen until it ruptures. It is very similar to the tensile testing experiment from AE 2610.
From the Test screen, select the red Browse methods
button on the right of the page.
Select the appropriate method
If using extensometer
Select AELabs>AE3610>AE3610 Project>AE3610TensileTest-Extens.im_tens
If using only strain gauges
Select AELabs>AE3610>AE3610 Project>AE3610TensileTest-Gauges.im_tens
Wait for the sample to initialize, and then select the red OK
button. The following method screen will open
Hit Balance all
on the bottom of the screen. All four of the live displays on the top of the screen should read zero or close to 0. If the strain starts drifting, ensure that the wires are connected to the adapter correctly and hit Balance all
again.
**Warning: The parameters of this method are based on the material properties and geometry of the Aluminum 6061 dogbone specimen from the AE 2610 Tensile Test lab. The following parameters may need to be altered to fit your test:
Test Stop Parameters - the parameters that will stop the test and stop loading on your specimen. The test is currently set to stop when the Aluminum 6061 dogbone specimen ruptures. These may need to be changed if are only loading the elastic region of your material, are not loading the specimen to rupture, or for another reason your group deems important.
Rate of Displacement - the displacement rate is how the Instron machine applies an increasing amount of load onto your specimen. It is currently set at 0.15 in/min but may need to be altered depending the material and geometry of your specimen.
Reach out to your TA if you feel these parameters need to be changed. DO NOT ALTER METHOD PARAMETERS WITHOUT THE HELP OF THE TAs OR THE LAB MANAGER**
Click Start
on the bottom right of the screen to begin the test. Observe the Force vs. Displacement chart to ensure the test is operating correctly.
If you feel you must stop the test at any point for any reason (e.g. you see yielding when you want to stay in the elastic region), you can hit the Stop
button on the bottom right of the screen.
You can export your data even if you hit the Stop
button during the test so you may use this as a method to Stop
the test and still export any data you have.
When the test is completed, either automatically due to Stop parameters or hitting the Stop
button, take a look at the final Force vs. Displacement chart to see if the test ran properly.
In the drop-down menu, select "Export to Tensile_1.csv"
In Windows Explorer, navigate to:
C:\Users\Public\Documents\Instron\Bluehill Universal\Templates\AE_Labs\AE3610\3610 Project\Fall 2022\Tensile.is_tens_Exports
In this folder should be a file labeled "Tensile_1.csv"
Open the file in Notepad and ensure the data is present and accurate.
If satisfied, rename this file with something meaningful (e.g. "A20_StainlessSteel_Rupture.csv")
Cut and paste this file to your team folder in:
C:\Users\Public\Documents\Instron\Bluehill Universal\Templates\AE_Labs\AE3610\3610 Project\Fall 2022\Group_Data
In the "Group_Data" folder, create a new folder with your group name if it does not already exist, and place the file there.
Close the Bluehill software
When prompted "Save changes to sample?", select No
.
Email the file to yourselves.
If the computer is not already connected to the internet, you must log into your eduroam to connect the computer to the WiFi. Remember to FORGET the network once you are done emailing the data to yourselves.
Turn off the Instron machine and Power down the computer.
One object of this experiment is to explore the displacement and strain behavior of structures using the digital image correlation technique. In addition, you will explore the interesting response of polypropylene, a material that exhibits different moduli in tension and compression. The experiment consists of two tests: 1) a four point bending test of a polypropylene specimen from which you will estimate the elastic moduli of polypropylene; and 2) a tension test of a second polypropylene specimen with a hole cut in it, from which you can determine the strain and stress concentrations caused by the hole. The loading for both tests will be accomplished using an Instron load frame.
In the context of structural testing, Digital Image Correlation (DIC) is a method for tracking the point-wise displacements of a structure (typically a surface of the structure) using a series of images of the structure undergoing deformation. DIC is a non-intrusive measurement technique since nothing has to be mounted to the specimen directly. Furthermore, DIC can measure real structural component geometries in real world conditions. The DIC measurements are primarily limited by image resolution, such that higher resolution images produce more accurate results. Alternatively, a higher-resolution displacement field can be captured by zooming the camera’s field of view to a smaller portion of the specimen of interest. To use DIC, it is usually necessary to prepare the specimen by painting a high-contrast speckle pattern on the surface so that subsequent pre- and post-deformation images can be analyzed to accurately determine the displacement field on the structural surface.
There are some requirements and rules-of-thumb for producing good speckle patterns. First, the pattern should be “random”; a highly organized and repeatable pattern would produce ambiguity in the displacement measurement. Second, the size of the speckles (or high contrast objects) should be roughly the same as (or slightly larger than) a 3x3 pixel region on the digital image for optimal tracking of the displacement. If the speckles are smaller than this, the image magnification can be changed to meet this criterion. Finally, the “density” of speckle features should be sufficient to have an average of 3-4 such features in a 10x10 pixel region. DIC does not rely on correlating a single speckle features, but rather multiple features to get an average displacement in a (multi-pixel) sub-region of the image.
DIC has its origin in speckle imaging approaches used in solid mechanics, and correlation-based analysis methods developed in the 1980’s for object tracking in image processing applications and particle-based velocimetry measurements in fluid mechanics. In fact, DIC is very similar to a common velocity-field measurement approach used in fluid mechanics, Particle Image Velocimetry (PIV). In both DIC and PIV, the individual displacements of many small subregions of an imaged area are obtained by comparing images before and after the displacement has occurred. For each subregion, the “before” (pre) and “after” (post) images are cross-correlated, sometimes using Fast Fourier Transform (FFT) algorithms. The displacement for that subregion is the one that provides the best correlation between the two images. In DIC, the displacement is the desired quantity. In PIV, this displacement is divided by the (short) time between the two images to obtain the local velocity.
This analysis process is typically performed after recording a sequence of images. In DIC, after the displacement field is calculated, the strain field can be determined. The two-dimensional surface displacement field is characterized as u(x,y) and v(x,y), where u and v are the displacements in the x and y directions for a point originally at location (x,y). With u and v determined, we can obtain the surface strain field using the strain-displacement relationships
(1)
Note that differentiating the displacement data amplifies the noise in the data; so advanced analysis software like the package used here employ additional processing approaches such as sophisticated smoothing to find the strain field from the displacement data.
In this lab, we will use the Aramis analysis software package.
You will also export the Aramis DIC data so you can analyze the full-field data.
Figure 1. The 3D DIC imaging systems, with two cameras mounted in a stereoscopic configurations and two light sources.
DIC (and PIV) systems come in different flavors. For example, the measured displacements of a thin region can be two-dimensional (2D) or three-dimensional (3D). In this lab, we will use a 3D DIC system (shown in Figure 1) that enables us to capture displacements in all three coordinate directions, including out-of-plane deformations. A 3D system provides more information than the more common 2D systems by adding an additional measurement. The 2D system requires only one camera (or equivalently, only camera imaging view point). To capture out-of-plane deflections, the 3D DIC system uses two cameras in a stereoscopic configuration. It is important to reiterate that 3D DIC (and PIV) systems provide three components of displacement (or velocity) from a surface (or thin planar region). There are also volumetric DIC and PIV approaches that provide 3D results for each location within a three-dimensional volume.
The four-point flexure or bending test is designed to test the flexural response of a slender beam. The goal of the four point bending test is to create a state of pure bending. Pure bending is a stress state where the bending moment is constant and the shear resultant is zero everywhere. The diagram on the left in Figure 2 illustrates a pure bending condition in which only opposing bending moments are applied to either end of the beam. Unfortunately, it is difficult to generate a pure bending moment in a real experiment. Instead, we will use opposing off-set point loads that generate a couple at either end of a beam. This configuration is shown on the right diagram in Figure 2. The advantage of this loading condition is that over the central span there are no shear loads and the beam is subject only to a bending moment.
Figure 2. Idealized pure bending load (left); four point bending test load (right).
Figure 3 illustrates the setup of the test apparatus for the four point bending test. The test specimen is placed on rollers which are placed below and above the specimen. The load frame applies a compressive load to the experimental apparatus which transmits point loads to the top and bottom of the beam. Rollers are used to ensure that simple support conditions are imposed.
Figure 3. Apparatus schematic for a four point bending test
Under ideal conditions in pure bending, the strain in the beam should be linear through the thickness:
(2)
Furthermore, if the material is linear elastic and isotropic, then Hooke's law applies and the bending moment can be calculated as follows:
(3)
where I is the second moment of area of the beam, which is given as
(4)
In Eq. (4), w is the width of the beam, and h is its depth. Therefore, if we impose M through a four point bending test, and can estimate from the digital image correlation results, we can infer the elastic modulus from
(5)
Polypropylene, however, exhibits different elastic moduli under tension and compression. As a result, the neutral surface is not at the geometric centroid of the cross-section and we have to use a composite beam analysis technique to find the elastic modulus under tension and compression. The measured strain will be offset from the geometric centroid, i.e., Eq. (2) has to be modified as follows:
(6)
where b is a strain offset.
(7)
Integrating Eq. (7) yields
(8)
(9)
Next, we know from equilibrium that there cannot be an internal axial load in the beam. Therefore, the axial resultant (N) must be zero:
(10)
(11)
(12)
Eq. (12) can be solved for the modulus under tension, giving
(13)
Inserting this into Eq. (11) provides an expression for the compressive modulus
(14)
Stresses around defects and sudden changes in a structure can be significantly higher than the average stress in the structure. These sharp increases in stress are called stress concentrations. A good example of a stress concentration is the behavior of the stress near a circular hole in a structure subject to uniform tension or compression. For an infinite plate loaded in-plane, the tangential stress around the edge of the hole has an analytic solution
(15)
(16)
(17)
In this lab, you will measure displacements and strains in a polypropylene test specimen (Figure 6) with a circular cutout subject to tension. The DIC system will enable us to visualize the distribution of strains around the hole and observe the stress and strain concentration.
Figure 4. Photograph of test specimen with hole.
If your specimen has a flat surface and good adhesion, self-adhesive labels with pre-printed speckle patterns can be directly applied to the specimen. The following process was followed to generate speckles for this lab:
Open Dymo Connect, set up the Extra Large shipping label in portrait, and import the image from the previous step, scaling it up to full size.
Print labels.
Remove any previous labels and/or any large residues or deposits on the specimen surface.
Degrease the specimen's surface.
Carefully apply the label in the region of interest without folding the edges, and applying firm pressure to aid its adhesion.
Use a sharp knife to trim away the excess label, taking care to not cut away any from the region of interest.
For certain surfaces, such as those that aren't flat, speckling paint directly onto the specimen is the preferred approach. One such paint application method is using a toothbrush as follows:
Lightly sand and deburr the specimen to remove any manufacturing artifacts. Wipe clean the specimen to remove any oils/residues that may prevent paint adhesion.
Apply a matte white paint base layer to the specimen to remove any reflectivity it inherently has. Do this using spray paint, shaking the can well and applying in multiple thin layers with ample time for each layer to dry before applying the next.
Dip a toothbrush into black paint, tapping off any excess. Having determined the appropriate distance away from the specimen to get the desired particle size beforehand, flick the brush's bristles to create a distribution of speckles over the base layer that will give good results. Consult the DIC manual for recommended particle size and coverage.
The 3D stereovision DIC system requires both of its cameras to be calibrated in order to project facets (i.e. speckles) observed in their own frames into real-world co-ordinates. To achieve this, an object with well-known geometry is placed at various positions and orientations in the field-of-view of the cameras, with images then taken for the software to perform a "stereo-triangulation".
Calibration can be a lengthy process, so please work quickly and diligently so as not to run out of time. If the cameras get knocked out of the position or get disorientation at any point during the labs, this calibration will need to be repeated before data can be acquired. Therefore, take great care to not bump or knock the cameras as you move around the Instron.
If you do need to recalibrate the cameras, the TAs will guide you through the process using the prompted procedure in the Aramis software.
Open the Aramis software from the Desktop
Click "File" and "New Project"
Select and arbitrary file name for this calibration (the file will not be saved)
Select the "3D" radio button
Click "Next" 3 times
Ensure the Max. deviation is set to 0.300 and then click "Next"
Click "Finish"
Enter calibration mode by clicking the black square button (4th from the left near the top of the screen)
Select the "CQG1171/55x44" option and click "Next"
Ensure the Focal length is 50.00 mm and click "Next"
Click Finish. You should now see a screen with two images (left image and right image).
Calibrate the system
With the help of the TAs, run through the 13 calibration images with the 55x44 calibration block and the white tilt stand.
It is best to rotate the Instron grips such that the front of the grips is facing the jackscrews when doing calibration
Use the slabs of metal and the instron grip rotation to raise and lower the calibration cube as necessary
For images taken without the tilt stand, ensure the image angles are within +/-0.3 degrees. For images taken with the tilt stand, ensure the images are within +/-5.5 degrees.
You can undo an image but you can only undo the current image. Once you click snap, the previous calibration image is saved and you cannot change it!
You CANNOT undo the last (13th) calibration image!!!!
This step may take some time! Be patient. You will have plenty of time to finish your lab but poor calibration will result in poor data!
Wait for the calibration result pop-up to appear. Ensure the deviations (both calibration and scale) are relatively small and then click OK.
Delete any previously taken images.
Enter Measurement mode by clicking the blue camera icon (5th from the left on the top of the screen).
Make sure the laser is off and the lights are on.
Take an unloaded reference image by clicking the blue camera icon on the bottom of the screen. This image will save as Stage 0.
Take a second unloaded image by clicking the same blue camera icon on the bottom of the screen. This image will save as Stage 1.
Exit measurement mode by clicking the blue camera icon (5th from the left on the top of the screen).
Mask the area of interest
Click on "Stage 0" in the explorer. Define a mask by clicking the blue and white "Define Mask" button on the top of the screen (6th from the left). A "define mask" pop up will show up.
Click "Mask all" (4th from the left). The bottom left screen should turn blue.
Click "unmask rectangle" (3rd from the right).
On the Stage 0 left image (bottom left image on the screen), left click and select the area of interest.
You should select a region such that you capture as much of the speckle pattern as possible and a small amount of dark space. Ask the TAs for guidance if necessary.
Dogbone specimen: You should capture the hole, speckle region, and a small amount of dark space on the left and right of the specimen.
4 pt. bending beam: You should capture the speckle region between the inner rollers, and a small amount of dark space above and below the specimen
You will see a green rectangle encompassing the region you select.
Right click inside the rectangle and the blue tint should be removed.
Click OK in the "Define Mask" pop-up and the pop-up will disappear.
With Stage 0 selected, click "Add a start point" (8th button from the left on the top of the screen)
Hold Control and click somewhere inside the green rectangle near the hole but not in the hole (for the open hole tension specimen) or inside the green rectangle on the speckle pattern (for the 4 point bending specimen).
In the pop-up, ensure that the intersection deviation is very low. Ask a TA if you are unsure.
Click Next. Click Create. Click Close.
Click "Compute Project" (3rd button from the right on the top of the screen).
Click Close once the project is computed and you will see the strain field.
Right click on the color legend, and under scaling select 2 sigma.
Observe the resulting surface plot and evaluate the quality of the surface detection. If you are satisfied with your results, move on – if not, assist the TA in re-calibrating the setup and repeat verification until you are satisfied.
Table 1. Four-point loading parameters to be used in experiment (defined in Figure 2)
Prepare the specimen:
Rotate the Instron grips such that it is facing forward if necessary. You will need 2 people to do this and make sure a TA is supervising this process!
Place the 4 point bending fixture into the instron grips. You will need 2 people to do this and make sure a TA is supervising this process!
Grip the bending fixture such that the grip section on the fixture is flush to the back of the clamps. If done correctly, there should be a gap in the front.
It should also be gripped ~3mm below where the grip starts to diverge. See the image below for reference.
Locate the beam specimen that has a self-adhesive speckle pattern label applied.
Measure the thickness and width of the beam specimen being careful not to touch the speckle pattern with your fingers
With the help of a TA, ensure that there is no folding, fading, or any other damage to the speckle pattern. If there is, apply a new speckle pattern sticker to the specimen.
With the help of a TA, position the beam in the test fixture using a 12-inch lower support length and a 4-inch upper support length.
Place the rollers in locations to produce the four-point bending parameters listed in Table 1. There are arrows on the fixture itself to depict where the rollers must be placed
Ensure the bottom rollers are on placed below the beam. These 2 rollers are labeled with a "B"
Ensure the top rollers are placed above the beam. These 2 rollers are labeled with a "T"
TA Note: If this line starts to become faded, redraw it with the sharpie in the green toolbox at the end of the lab.
Have a TA verify the position of the beam by turning on the laser, and using a slip of paper placed flush to the front surface of the beam. If the laser is in the center of the image on the Aramis software, then the beam placement is accurate.
Make sure to turn the laser off.
Zero the load and displacement on the Instron software.
Calibrate the system:
If you have not yet calibrated the cameras, follow the calibration procedure above to calibrate the DIC camera system.
If you have already calibrated the system, you may skip this step.
Verify the camera calibration:
Follow the camera calibration verification procedure above.
Perform the experiment:
Delete any previously-taken images
Enter Measurement mode by clicking the blue camera icon (5th from the left on the top of the screen).
Make sure the laser is off and the lights are on.
Take an unloaded reference image by clicking the blue camera icon on the bottom of the screen. This image will save as Stage 0.
Jog the Instron down until you've applied a bending load of 150 lbf.
Take an image of the specimen under load. This image will save as Stage 1.
Immediately unload the specimen, noting that the beam will slowly spring back to its natural unloaded position meaning you will need to jog the Instron further than you initially realize.
Exit measurement mode by clicking the blue camera icon (5th from the left on the top of the screen).
Mask the area of interest
Click on "Stage 0" in the explorer. Define a mask by clicking the blue and white "Define Mask" button on the top of the screen (6th from the left). A "define mask" pop up will show up.
Click "Mask all" (4th from the left). The bottom left screen should turn blue.
Click "unmask rectangle" (3rd from the right).
On the Stage 0 left image (bottom left image on the screen), left click and select the area of interest.
You should select a region such that you capture as much of the speckle pattern as possible and a small amount of dark space. Ask the TAs for guidance if necessary.
4 pt. bending beam: You should capture the speckle region between the inner rollers, and a small amount of dark space above and below the specimen
You will see a green rectangle encompassing the region you select.
Right click inside the rectangle and the blue tint should be removed.
Click OK in the "Define Mask" pop-up and the pop-up will disappear.
With Stage 0 selected, click "Add a start point" (8th button from the left on the top of the screen)
Hold Control and click somewhere inside the green rectangle on the speckle pattern.
In the pop-up, ensure that the intersection deviation is very low. Ask a TA if you are unsure.
Click Next. Click Create. Click Close.
Click "Compute Project" (3rd button from the right on the top of the screen).
Click Close once the project is computed and you will see the strain field.
Right click on the color legend, and under scaling select 2 sigma.
Inspect the results for what you would expect to see for this load case. If you do not attain a usable result, try to remedy the situation in this order:
Re-process the existing images with a different start point
Verify lighting, focus, shutter time, with the TAs and re-capture unloaded and loaded images before re-processing
Re-apply a speckle label and re-process
Re-calibrate the cameras and re-process
If you are satisfied with your results, remove the specimen from the Instron and export your data.
Click Export > Export All Points > Select "AE3610_01082024_Spring2024_ExportConfig", select your filename and save location, then hit OK.
Make sure the "Selected Points Only" check box is unchecked before exporting data and the proper config file is used!
Prepare the specimen:
Rotate the Instron grips such that it is facing sideways if necessary. You will need 2 people to do this and make sure a TA is supervising this process!
Locate the open-hole dog-bone specimen that has a self-adhesive speckle pattern label applied.
Measure the specimen's thickness, width, and hole diameter.
With the help of a TA, ensure that there is no folding, fading, or any other damage to the speckle pattern. If there is, apply a new speckle pattern sticker to the specimen.
With the help of a TA, position the specimen in the clamping test fixture.
Zero the load and displacement on the Instron software.
Calibrate the system:
If you have not yet calibrated the cameras, follow the calibration procedure above to calibrate the DIC camera system.
If you have already calibrated the system, you may skip this step.
Verify the camera calibration:
Follow the camera calibration verification procedure above.
Perform the experiment:
Delete any previously-taken images
Enter Measurement mode by clicking the blue camera icon (5th from the left on the top of the screen).
Make sure the laser is off and the lights are on.
Take an unloaded reference image by clicking the blue camera icon on the bottom of the screen. This image will save as Stage 0.
Jog the Instron up until you've applied a tensile load of 2.2 kN.
Take an image of the specimen under load. This image will save as Stage 1.
Immediately unload the specimen, noting that an unloaded specimen has a load reading of 0 kN. You may ungrip one side of the specimen to fully unload the specimen.
Exit measurement mode by clicking the blue camera icon (5th from the left on the top of the screen).
Mask the area of interest
Click on "Stage 0" in the explorer. Define a mask by clicking the blue and white "Define Mask" button on the top of the screen (6th from the left). A "define mask" pop up will show up.
Click "Mask all" (4th from the left). The bottom left screen should turn blue.
Click "unmask rectangle" (3rd from the right).
On the Stage 0 left image (bottom left image on the screen), left click and select the area of interest.
You should select a region such that you capture as much of the speckle pattern as possible and a small amount of dark space. Ask the TAs for guidance if necessary.
Dogbone specimen: You should capture the hole, speckle region, and a small amount of dark space on the left and right of the specimen.
You will see a green rectangle encompassing the region you select.
Right click inside the rectangle and the blue tint should be removed.
Click OK in the "Define Mask" pop-up and the pop-up will disappear.
With Stage 0 selected, click "Add a start point" (8th button from the left on the top of the screen)
Hold Control and click somewhere inside the green rectangle near the hole but not in the hole.
In the pop-up, ensure that the intersection deviation is very low. Ask a TA if you are unsure.
Click Next. Click Create. Click Close.
Click "Compute Project" (3rd button from the right on the top of the screen).
Click Close once the project is computed and you will see the strain field.
Right click on the color legend, and under scaling select 2 sigma.
Inspect the results for what you would expect to see for this load case. If you do not attain a usable result, try to remedy the situation in this order:
Re-process the existing images with a different start point
Verify lighting, focus, shutter time, with the TAs and re-capture unloaded and loaded images before re-processing
Re-apply a speckle label and re-process
Re-calibrate the cameras and re-process
If you are satisfied with your results, remove the specimen from the Instron and export your data.
Click Export > Export All Points > Select "AE3610_01082024_Spring2024_ExportConfig", select your filename and save location, then hit OK.
Make sure the "Selected Points Only" check box is unchecked before exporting data and the proper config file is used!
Thickness and width of the beam, and load value used in the four-point bending test.
Displacement and strain data from the Aramis software for the four-point bending test.
Thickness and width of the specimen, hole diameter, and load value used in the open-hole tension test.
Displacement and strain data from the Aramis software for the open-hole tension test.
For the open-hole tension test, compute the stresses in your specimen based on the measured strains and the appropriate measured modulus of elasticity for polypropylene.
Table of beam dimensions and load value used in the four-point bending test.
Table of specimen dimensions, hole size and load value used in the open-hole tension test.
For the four-point bending test, plots of the axial and transverse normal strain fields as a function of y (with the y-axis as defined in Figure 5), at three axial locations between the two inner load points (i.e., within the constant moment region). One of these should be the center axial location.
Figure 5. Through-thickness stress-strain distribution for material with different moduli in compression and tension.
A single graph containing plots of the tangential stress along a line perpendicular to the length dimension of your specimen that passes through the center of the hole, for each loading.
A single graph containing plots for each of the loadings of the tangential stress (normalized by its value far from the hole) as a function of angle around the hole, close to the edge of hole.
Apply the following settings before exporting to PDF (the geometry can change, but this is set up specifically for Dymo Extra Large shipping labels using a LabelWrite 4XL thermal printer):
Ensure that the front of the fixture is facing forward--The side labeled "Front"
Ensure the letter "F" on all 4 rollers is facing the cameras. Furthermore, ensure all 4 rollers are flush to the front surface of the fixture. See the image below
Place the beam on the rollers such that the front face of the beam is resting just behind the line drawn on the rollers. Use the lines on the bottom rollers to do this. See the image below
Pressure Tap | Corresponding Tap Number on VDAS |
---|---|
In the VDAS software, click the "Start timed acquisition of data" button
Export this data to a .xlsx file by clicking the "Export recorded data to Excel XLSX file" button
In the VDAS software, click the "Start timed acquisition of data" button
Export this data to a .xlsx file by clicking the "Export recorded data to Excel XLSX file" button
Connect the left-hand strain gauge to the left-hand strain gauge adapter module (120 Ohm A) and the right-hand strain gauge to the right-hand strain gauge adapter module (120 Ohm B). In each case, connect Port 1 of the adapter module to one side of the strain gauge, Port 2 to the other side of the strain gauge, and install a jumper wire between Ports 2 and 3. See the image below for a reference. Gently tug on each wire to ensure a good secure connection is made.
Click on the System Details tab in the top right corner of the screen
Open the Strain Gauge Transducer A settings by clicking the button under System Settings. Ensure the following settings:
Repeat the above step 3 for Strain Gauge Transducer B
Run the VI by pressing the button, creating a folder for your group and setting that as your save folder, before verifying successful hardware connection (no flags thrown on run)
Below your BOM table, include URLs for your Home Depot and McMaster shopping carts ()
A GTAE Reponsible Itemized Bill of Materials for all items you will use from the GTAE communal stock, including information on:
GTAE will also provide any consumables that benefit AE3610 as a whole. This includes fluids-specific and structures-specific consumables as well as general consumable items. All provided consumables are outlined below under . Note that general consumables are provided for typical small amounts of use. If you need a gallon of epoxy to mold something, that is not a typical small use! In this case, you must include epoxy in your "Team Responsible BOM" and add the cost of this item to your total costs. We will make general consumables available to you in a common location for use (ESM G2) during the rest of the semester. You must still include small amount consumables provided by GTAE in your GTAE Responsible BOM list! This is to ensure enough of everything gets purchased.
Click on the top right, then
Insert YOUR OWN email in the 'To' field, ensure 'Email me a copy' is checked, and hit
Click on the top right of the page
Click located roughly in the center of the page, to the left of the 'Your Order' section
Click , inputting YOUR OWN email address and hitting 'Send'
Find the email from Home Depot in your email account and right-click on the button before selecting 'Copy link address' (for Chrome on W10, other browsers will vary)
Within GT-AE, the and its mentors are available to you. If any of your group have completed Machine Shop training, you also have access to the machines therein. Some limited machine shop machining assistance *may* be provided by way of the Lab Manager and TAs, for water-jetting metals and simple machining. All Machine Shop use needs to be arranged via your assigned TA; the Machine Shop machinists SHOULD NOT BE APPROACHED FOR INDIVIDUAL HELP.
Outside of GT-AE, GT's other maker spaces are also available for your use, for example the and .
The
A has been created for you to use to mount your wing onto the load-measurement turntable. This mount is described in detail under "Wing Mount Description and Dimensions".
, including:
Handheld
(for Dogbone Clamp Fixture)
(for 4-Point-Bending Fixture)
(for V-Groove Fixture)
In each case, connect Port 1 of the adapter module to one side of the strain gauge, Port 2 to the other side of the strain gauge, and install a jumper wire between Ports 2 and 3. Gently tug on each wire to ensure a good secure connection is made.
If you are satisfied with the data, Click the Export button in the top right of the screen
Our 3D DIC system, which employs two 5-megapixel cameras, will measure the displacements within a test volume that is centered on the test specimen. While the 3D DIC system provides accurate 3D displacement fields, it requires additional calibration effort compared to a simple, single-camera 2D system. The cameras are calibrated by orienting a thermally balanced plate with calibrated markings on it within the test volume.
The Aramis software will guide you through the calibration.
The through-thickness stress distribution is shown in Figure 5. The y-location of the neutral surface can be found as . The bending moment can be found by a similar integration as that in Eq. (3), except using Eq. (6) for the strain and integrating separately on either side of the neutral surface, i.e.,
Substituting our expression for in (8) and simplifying gives
Integrating Eq. (10) and replacing with yields
Now combining Eq. (11) and Eq. (9) for the bending moment, eliminating __ and using our expression for the 2nd moment of area, Eq. (4), results in
where the coordinate θ runs circumferentially around the circular hole and is the average stress in the specimen far away from the hole. The direction θ=0 is aligned with the loading direction, or the x-axis in our geometry. The maximum value of the stress (due to the hole) normalized by is known as the stress concentration factor.
The stresses in planar cylindrical coordinates (i.e., , , and ) can be calculated from Cartesian stresses (i.e., , and ) using standard coordinate transformations, for example
A critically important step for a DIC experiment is to prepare the specimen's surface and to apply a speckle pattern that the software can interpret and transform into a strain field. Various resources such as and detail how best to do this. For this lab, the specimens have already been prepared, so the steps below are for your enlightenment.
Download the software.
Apply the following settings before exporting to PDF (the geometry can change, but this is set up specifically for Dymo Extra Large shipping labels using a LabelWrite 4XL thermal printer):
Convert the PDF to a high-resolution image using or similar. This file can also be found below this list.
Ensure that the front of the fixture is facing forward--The side labeled "Front"
Ensure the letter "F" on all 4 rollers is facing the cameras. Furthermore, ensure all 4 rollers are flush to the front surface of the fixture. See the image below
Place the beam on the rollers such that the front face of the beam is resting just behind the line drawn on the rollers. Use the lines on the bottom rollers to do this. See the image below
With the specimen under load, inspect the speckle label for any signs of peeling/buckling/distortion. If you observe any, you will need to unload the specimen and re-apply a label before following all previous steps to get back to this step.
With the specimen under load, inspect the speckle label for any signs of peeling/buckling/distortion. If you observe any, you will need to unload the specimen and re-apply a label before following all previous steps to get back to this step.
Using a best linear fit for the appropriate region of your plots from the four-point bending test, compute the slope () and intercept (b) relative to a coordinate axis centered on your specimen.
From your experimental data, determine the tensile and compressive moduli ( and ) for polypropylene.
From your plots of transverse normal strain divided by the axial strain , determine a value for Poisson’s ratio () for polypropylene.
“Image graphs” or “full-field color plots” of the axial (horizontal) and transverse (vertical) displacement field for the field of view analyzed by the Aramis software in the bending test. Choose your color scaling wisely to accentuate any important gradients. Be sure to also include color bars showing how your colors map to the values of displacement.
Here an “image” means a false-color image, where each displacement (or strain) measurement location is like a pixel in the false-color image, and the color of the pixel corresponds to the value being shown (axial displacement in this case). You can do this, for example, with the image function in Matlab
Images of , and for the field of view imaged and analyzed by the Aramis software in the bending test, with color bars.
Plot of the transverse normal strain divided by the axial strain along the center line used for Result 5.
Table of measured values of , b, and the polypropylene properties , , and . Include in your table the published values for the polypropylene properties (include your source for those values).
Images of the axial strain () and transverse strain () fields for the region analyzed by the Aramis software in the open-hole tension test, with color bars, for both loadings.
Images of the axial () and transverse stress () fields based on your measured strains and measured modulus of polypropylene, with color bars, for both loadings.
Images of the normal radial stress () and normal tangential stress () fields for the open-hole tension tests, with color bars, for both loadings.
Material
**** °C (°F)
ANSI Type
Allowable atmosphere
Avg Output mV/100°C
Tungsten/ tungsten 26% rhenium
2320/4210
-
inert, (nonoxidizing)
1.7
Tungsten 5% rhenium/ tungsten 26% rhenium
2320/4210
-
inert, (nonoxidizing)
1.6
Platinum 30% rhodium/ platinum 6% rhodium
1820/3310
B
oxidizing, inert
0.76
Platinum 13% rhodium/ platinum
1770/3200
R
oxidizing, inert
1.2
Platinum 10% rhodium/ platinum
1770/3200
S
oxidizing, inert
1.0
Chromel/alumel
1370/2500
K
oxidizing, inert*
3.9
Chromel/constantan
1000/1830
E
oxidizing, inert*
6.8
Iron/constantan
1200/2193
J
reducing, inert, vacuum
5.5
Copper/constantan
400/750
T
mild oxidizing, reducing vacuum, inert
4.0
Temperature Range
0-500 °C
500-1372 °C
Voltage Range
0-20,644 μV
20,644-54,886 μV
0.000000
-1.318058×10
2.508355×10
4.830222×10
7.860106×10
-1.646031×10
-2.503131×10
5.464731×10
8.315270×10
-9.650715×10
-1.228034×10
8.802193×10
9.804036×10
-3.110810×10
-4.413030×10
-
1.057734×10
-
-1.052755×10
-
Voltage (Volts)
Flowrate (cc/min)
1.7
175
2.2
227
2.8
255
3.0
300
Value
a (in.)
4
s (in.)
4
L (in.)
12
Working Section Static Tap 1-25 in order
Taps 1-25 in order
Airfoil Top Surface Pressure Taps 1 and 2
Tap 26 and 27 (not necessarily in order)
Static Pressure at Contraction Cone
Tap 28
Long Column
Short Column
Length (in.)
29
24
Width (in.)
0.75
0.75
Thickness (in.)
0.50
0.50
Value |
a (in.) | 4 |
s (in.) | 4 |
L (in.) | 12 |