You are on page 1of 3

1

Measurement of Youngs Modulus Using Strain Gauges (December 2011)


Gustavo Guillen, Student, University of Idaho
AbstractIn todays safety-conscious society, product liability is more important than ever before. One of the most significant measures of safety in many modern materials is Youngs Modulus. The purpose of this experiment was to substantiate the value of Youngs Modulus of Elasticity commonly used for commercial steel. In order to do so, we used a resistance type strain gauge to measure strain in a loaded cantilever beam. The two gauges that were bonded to the beam were then connected to a Wheatstone bridge circuit, and finally an amplifier. The beam was loaded and unloaded numerous times in order to gather multiple data points to assure accurate results. Meanwhile, the stress was calculated and related to the strain measured in the cantilever. These two values were then plotted and fitted using a linear regression curve; the slope of this curve is Youngs Modulus of Elasticity. Unfortunately, our results did not correspond to the published values of Youngs Modulus, possibly due to a data acquisition or systematic error, so we were left to work the calculations with a fellow colleagues data set. Index TermsYoungs Modulus of Elasticity, Cantilever Beam, Strain Gauge, Half-Arm Wheatstone Bridge Circuit

specimen of interest, k is the materials spring constant, and x is the displacement from its equilibrium position. After Hookes discovery, many scientists attempted to apply his law to their own areas of study. Sometime later in 1807, a scientist by the name of Thomas Young discovered that materials display very similar traits when the stress applied to them is related to the strain. The most notable thing he discovered is that this relationship is linear in the elastic region, and nonlinear in the inelastic region. This property was evaluated and given the symbol E, for Youngs Modulus, also called Modulus of Elasticity. II. ANALYSIS As mentioned before, a material that behaves elastically (returns to its original shape after being deformed) can be described by Equation (2). After some algebraic manipulation, it can be shown that:

(3)
Where is the normal stress, or normal force divided by the cross sectional area of the specimen. E is the Modulus of Elasticity, and is the strain as described in Eq. (1) [2]. For our experiment, in order to calculate the stress due to the load at the end of the cantilever, we must use the following equation:

I. INTRODUCTION

a force acts on a body, it always causes a deformation. Whether the deformation is as apparent as stretching a rubber band, or as minute as pressing against a steel plate, it is always there. For all practical purposes, this deformation is called strain [1]. More specifically, strain () is the change in length over the original length of the specimen:
HEN

(4)
Where M is the bending moment at the point of interest (the end, in our case), y is the distance from the neutral axis, and I is the moment of inertia based on the geometry (a rectangle for our beam). The setup for this experiment can be seen in Figure 1 below [4]. One gauge actively measures the tensile strain [top] and the other measures the compressive strain [bottom]: Strain Gauges

(1)
As we can see, strain is dimensionless since it is simply a ratio of length. Moreover, this value is typically very small, no more than the order of a thousandth. In 1635, Robert Hooke was born in England, where he later became a prolific architect, philosopher, and mathematician. His contributions to mechanics are still widely used today. In 1660, he discovered the law of elasticity, aptly named Hookes Law. Hookes Law states that, within certain limits, the stress in a material is proportional to the strain which produced it [2]. It can be represented mathematically by the following equation:

(2)
Where F is the force exerted at the end of the spring, or a
Gustavo Guillen is with the Mechanical Engineering Department, University of Idaho, Moscow, ID 83843 USA (e-mail: guil0470@vandals.uidaho.edu).

A. Fig. 1 End loaded cantilever beam with bonded strain gauges.

From Fig. 1 above, x is the distance from the load at the end of the beam. Additionally, notice the dimensions described to the

2 right. These dimensions are used to calculate the Moment of Inertia, I, used in our stress calculation in Eq. (4). As for the strain gauges, the experiment called for two metallic gauges with uniform cross section. Knowing this information, we can find the resistance of a conductor (the gauge) with the following equation [5]: III. EXPERIMENTAL PROCEDURE The first step of the procedure is to make a Simulink program in Matlab that properly acquires the output voltage from the circuit. Measurements of the beam itself were taken, as well as their associated uncertainties. These will later be used to calculate E. From here, the outputs from the bridge circuit are the inputs of the op-amp. Since the output values from the circuit will be so small, a gain, Ga , of more than 400 is required. Loads are ready to be applied to the beam: we first made a voltage measurement with the beam unloaded and alternately loaded and unloaded the beam with weights from 100 g to 2kg. With each load, a loaded/unloaded voltage measurement was taken. The signal from the Wheatstone bridge went through the resistance from the gauges (GF = 2.012), as well as two additional resistors with a value of R= 350 k. The output voltage from the circuit was then passed through an AD622 op-amp that was connected to an external resistance of Rg= 100.6 k. Using this resistance in Eq. (7) results in a gain of Ga= 502.988. Finally, the output of the op-amp was passed through an NI USB-6008 data acquisition device, and thus into our Simulink program. All of the data was carefully input into a Matlab m-file that processed the data and produced a plot of vs. . From Eq. (3), we can see that by using a linear regression curve, we can find the Modulus of Elasticity (the slope of vs. ). IV. RESULTS Figure 3 shows a plot of stress vs. strain for a steel cantilever beam. An important disclaimer should be noted: the results shown are not from our experiment, but from a fellow colleague, Zane K. An explanation if this is to follow in section V. Also note the size of the plot. A 3.5 inch TIFF image would not allow proper viewing of the axis, so a larger format had to be used. Since this experiment was to simply derive Youngs Modulus for steel, only the linear portion of the stress-strain curve is shown. Figure 3 very clearly shows this relationship. It is important to realize that the stress-strain graph is not entirely linearif we were to begin plastically deforming the beam, our graph would begin to exhibit a non-linear relationship between stress and strain. Using a linear regression of the data, we can show the resulting equation shown on the top-left of Figure 3. Note the resulting slope of the equation; , or 306 psi, which is the widely-used value for the Modulus of Elasticity for steel. To put into perspective, Youngs Modulus for rubber is anywhere from .16 to .66 psi, depending on its composition. Using Eq. (8), we can solve for the standard error of the fit for our linear regression curve. The norm of the residuals for our fit resulted in a value of norm= 338.16. Using this, in conjunction with N= 20 data points, and order m= 2 of the polynomial used, we get Syx= 82.016 psi, which is acceptable considering our objective value is on the order of 106 psi.

(5)
Where R is the resistance in the strain gauge, Lw is the length of the wire, and e is the resistance of the wire material. When the gauge is deformed, it produces a change in resistance. This change can be represented by the gauge factor, or GF, of the strain gauge, which was given in our experiment. Typical values of GF are around two. Since the voltage change is usually very small, a half-arm bridge circuit was used to acquire this small change. Knowing the value of a given resistor, we can find the gain of the circuit as [5]:

(6)
The following figure represents our Wheatstone bridge circuit setup, where R#1 and R#2 are the resistance values of the respective strain gauges [4]:

Fig. 2 Cantilever beam with bonded strain gauges with Wheatstone Bridge Circuit shown to the right.

The significance of this setup is the output voltage that results when an input voltage is supplied. Unfortunately, the magnitude of the output voltage is very low, so the signal must be passed through an operational amplifier. Doing so will allow us to amplify the magnitude of the signal. However, we must take this amplification, or gain, into account. We can do so by the following equation:

(7)
After proper setup of the circuit and testing apparatus (cantilever beam), data points can start being taken (explained later in the paper). Obviously, through taking measurements, errors can occur, thus propagating though the process resulting in faulty results. To estimate this uncertainty, in Youngs Modulus in our case, we can evaluate the standard error of the linear regression curve, or the standard error of the fit [5]:
( )

(8)

Where norm is the norm of the residuals and ( ), where N is the number of data points and m is the order of the polynomial of the regression curve.

Fig. 3 Loading results after taking 20 data points. The resulting slope of the fitted line is E, Youngs Modulus.

REFERENCES V. CONCLUSION As previously mentioned, the results above were not created using our original data, but from a colleagues data. When we ran our Matlab program with the data we collected, we achieved very consistent, systematic data, however our resulting E was 1.311 psi, clearly orders of magnitude off of the target value. The errors could have possibly been due to a data collection error, a systematic error, or possibly even a circuit error (even though no mistakes were found when the circuit was peer reviewed). Our data suggests that there was absolutely no linearity error, since all of our data points laid directly on our regression line. And finally, the error could have not been caused from our Matlab program since it processed a correct result using someone elses clean data. Unfortunately, the source was never found. Although using our colleagues data, we got very clean results. The data very clearly illustrates the linear relationship of the stress-strain curve in the elastic region. Furthermore, the resulting E corresponds nicely with published values of E for steel. Keep in mind that when the material exceeds this elastic region, the experimental results are no longer valid for describing Youngs modulus. Even though the effects of axial and torsional strain are very minute and most often negligible in experiments such as these, they were obviously not accounted for. While some similar configuration types can be used to measure torsional strain, NIs software scaling does not support these configuration types [6].
[1] Jacob Lubliner. University of California at Berkley (Revised ed.) [Online]. Available: http://www.ce.berkeley.edu/~coby/plas/pdf/book.pdf Mischke Shigley. Intermediate Mechanics of Materials for Component Design. McGraw Hill, 1989. James M. Gere and Barry J. Goodno, Mechanics of Materials. Cengage Learning (Seventh Ed.), Mason, OH, 2009. Wolbrecht, Eric. Experiment 7: Measurement of Youngs Modulus using Strain Gauges. 2007. Blackboard [Online]. Available: https://blackboard.uidaho.edu/webct/urw/tp0.lc5116001/cobaltMainFra me.dowebct R. S. Figiola and D. E. Beasley, Theory and Design for Mechanical Measurements. Wiley, 2006 National Instruments, Strain Gauge Configuration Types, NI Developer Zone [Online] Available: http://zone.ni.com/devzone/cda/tut/p/id/4172

[2] [3] [4]

[5] [6]

You might also like