You are on page 1of 182

Physics 135AL

Physics for the Life Sciences

Laboratory Manual
Spring 2007

Department of Physics and Astronomy University of Southern California Los Angeles CA 90089-0484

ii

Acknowledgments The current edition of this laboratory manual was written by Dr. Gkhan Esirgen. The following people contributed to past editions: Professor Robert Cole (emeritus), Ty Buxman, Kristin Sabo, and Robert Knol.

iii

Contents Introduction Experiment 1 Experiment 2 Experiment 3 Experiment 4 Experiment 5 Experiment 6 Experiment 7 Experiment 8 Experiment 9 Experiment 10 Experiment 11 Experiment 12 Appendix A Appendix B Physics 135AL laboratory rules and policies Data analysis: errors and graphing Data analysis: application Two-dimensional motion Circular motion Collisions in one dimension Rotational kinematics Forces and torques in static equilibrium I: cantilever Elasticity: stress and strain Fluid flow Sound Ideal gas and the absolute zero of temperature Thermal conduction and radiation: case study and Leslies cube DataStudio starter manual Error analysis

iv

Name: Partners name: Date and time:

PHYS135AL

EXPERIMENT I DATA ANALYSIS: ERRORS AND GRAPHING

1.0 Objective
You will learn how to determine the most common sources of error in experimental data and how to graphically represent data and use it for further analysis. Then you will study position, velocity, and acceleration data and graphs.

2.0 Recording data


2.1 Units
In most sciences the common unit system in use is the metric MKS (meters, kilograms, seconds) system. Less common is the metric CGS (centimeters, grams, seconds) system, and then finally the old standard, the English system (feet, slugs, seconds). You should have previously encountered these various systems in your studies so there wont be a discussion of them here. Simply realize that combining values that have either been converted to, or were already in the same unit system is extremely important because unit definitions for more complex quantities typically depend upon the specific system. (The group which seems to use the most contradictions to this rule are astronomers, who love to mix and match unit systems in odd ways. It is true that all fields develop their own jargon in this sense, but in physics lab were going to remain purists. In a sense this is what physics is) Recording units with data is as important as the numeric value itself, if not more so. Every time you record data or the results of a calculation, you must include its units. 2.1.1 Why? Give two reasons.

2.2 Scientific notation


Scientific notation is simply shorthand for recording large or small values, or values with a large number of placeholders. The Milky Way galaxy is believed to be 100,000 light years (Ly) across. In scientific notation this becomes 1 x 105, which translates to a leading 1 followed by 5 zeroes. If the exponent on the 10 were negative, the 1 would have been preceded by 4 zeroes, i.e., .00001. (The decimal was moved 5 places to the left rather than the right.)

2.3 Significant figures


It is important to be honest when reporting a measurement so that it does not appear to be more accurate than the equipment used to make the measurement allows. We can achieve this by controlling the number of digits, or, significant figures, used to report the measurement.

EXP I

Determining the number of significant figures The number of significant figures in a measurement, such as 2.531, is equal to the number of digits that are known with some degree of confidence (2, 5, and 3) plus the last digit (1), which is an estimate or approximation. As we improve the sensitivity of the equipment used to make a measurement, the number of significant figures increases.

Postage Scale Two-pan balance Analytical balance

31g 2.53 0.01 g 2.531 0.001 g

1 significant figure 3 significant figures 4 significant figures

Rules for counting significant figures are summarized below: Zeros within a number are always significant. Both 4308 and 40.05 contain four significant figures. Zeros that do nothing but set the decimal point are not significant. Thus 470,000 has two significant figures. Trailing zeros that arent needed to hold the decimal point are significant. For example 4.00 has three significant figures. If you are not sure whether a digit is significant, assume that it isnt. For example if the directions for an experiment read Put the lens at 60 cm, assume that the distance is given to one significant figure, but if 60. cm, then there are two, likewise if 60.0 cm, then there are three significant figures. Addition and subtraction with significant figures When combining measurements with different degrees of accuracy and precision, the accuracy of the final answer can be no greater than the least accurate measurement. This principle can be translated into a simple rule for addition and subtraction: When measurements are added or subtracted, the answer can contain no more decimal places than the least accurate measurement.

150.0 g dynamics cart + 2.531 g penny 152.5 g total mass

(one significant figure after the decimal point) (three significant figures after the decimal point) (one significant figure after the decimal point)

Multiplication and division with significant figures The same principle governs the use of significant figures in multiplication and division: the final result can be no more accurate than the least accurate measurement. In this case, however, we count the significant figures in each measurement, not the number of decimal places: when measurements are multiplied or divided, the answer can contain no more significant figures than the least accurate measurement.

EXP I

Example: To illustrate this rule, let's calculate the cost of the copper in an old penny that is of pure copper. Let's assume that the penny has a mass of 2.531 grams, that it is essentially of pure copper, and that the price of copper is 67 cent per pound. We can start converting from grams to pounds.
1 lb 2.531 g ----------------- = 0.005580 lb 453.6 g

We then use the price of a pound of copper to calculate the cost of the copper metal.
67 cent 0.005580 lb ----------------- = 0.37 cent 1 lb

There are four significant figures in both the mass of the penny (2.531) and the number of grams in a pound (453.6). But there are only two significant figures in the price of copper; so the final answer can only have two significant figures. Now multiply the numbers 125.75 cm x 91 cm x 100.0 cm and answer the following questions. 2.3.1 Which one of the three values has the smallest number of significant figures?

2.3.2 Do the multiplication. What are the units of your answer?

2.3.3 How many significant figures should your answer have?

2.3.4 And the correct answer is?

Rounding Off When the answer to a calculation contains too many significant figures, it must be rounded off. There are 10 digits that can occur in the last decimal place in a calculation. One way of rounding off involves underestimating the answer for five of these digits (0, 1, 2, 3, and 4) and overestimating the answer for the other five (5, 6, 7, 8, and 9). This approach to rounding off is summarized as follows. If the digit is smaller than 5, drop this digit and leave the remaining number unchanged. Thus, 1.684 becomes 1.68. If the digit is 5 or larger, drop this digit and add 1 to the preceding digit. Thus, 1.247 becomes 1.25. But remember to round off only your final answer. If you round off during intermediate steps, accuracy of your final answer might be lost.

EXP I

Also use scientific notation whenever appropriate, keeping the significant figures. For example if there are four significant figures in the calculated result 103450 ft, then it should be expressed as 1.035 x 105 ft in scientific notation.

3.0 Error analysis


Most lab situations concern themselves with the taking, processing, and analysis of data from which theory is derived, supported, or disproved. It is always wise to understand the limits inherent in your numbers, for if there is one thing that is certain in this universe, it is that nothing is exact. So if nothing is exact, just how close to exact is your data?

3.1 Instrument errors


All measuring tools have error associated with them, usually known as the instruments precision. In principle when you use an instrument such as a ruler, your least significant figure (the rightmost digit) should only be an estimate, and the other digits must be directly inferable from the instrument. For example with a ruler with mm spacings, a measurement might be expressed in centimeters as 14.12 cm, where 14.1 cm is directly read off the mm division marks and 0.2 mm is estimated. 3.1.1 Using a meterstick measure the width of your lab table and record it here.

3.1.2 What is the smallest unit on your meter stick (mm, cm, or m)?

3.1.3 Can you say that the measurement of the tables width is closer to one (of the smallest) marks than another?

The smallest increment you can read on your instrument is known as the least count of that instrument. For the meter stick, the least count is the value recorded in 3.1.2. The definition of instrumental uncertainty for an instrument is simply 1/2 x least count of that instrument. 3.1.4 What is the metersticks measurement uncertainty? Does this make sense?

Anyone else should be able to look at data youve taken and without knowing details of the experiment understand how exact it is and what the units are. The uncertainty Xi must always be recorded with the value Xi itself, and the units must be included:

X i X i (unit)

(EQ 1)

EXP I

3.1.5 In the correct format, write down the width of the table and its uncertainty. Remember the units. Note that since this format involves the addition/subtraction operation, the number of significant figures after the decimal point should be the same for your measurement and uncertainty; otherwise, your expression doesnt make sense.

3.2 Fractional and percent deviation


The fractional deviation answers the question What fraction of the value is the uncertainty? and is given by X/X. The percent error is simply 100 X/X.

3.3 Propagation of error through calculations


In general errors are of an additive nature, but to simply add them overestimates the deviation (see a stats book for details). As with significant figures, there are rules for addition/subtraction, multiplication/division, and here we add a rule for powers. These rules are presented without proof for everyones collective sanity. Addition and Subtraction: Let C=A+B or C=A-B. Then the square of the error after addition or subtraction in both cases is the sum of the squares of the input errors. Mathematically, we write

C =

( A ) + ( B )

(EQ 2)

Multiplication and Division: Let C = A x B or C=A/B. The key to errors in multiplication and division is to look at the fractional errors:

C = -----C

2 2 A + B ---------- B A

(EQ 3)

Powers: Let A = Bn, then the fractional error of A is given as:

A = n B ----------B A
3.3.1 Calculate the error when you subtract 36.0 +/-.2 m from 18.3 +/-.7 m

(EQ 4)

EXP I

3.3.2 The volume of a cylinder is V = r2h. Write down the correct general formula to determine the error in the volume of the cylinder, V, given r and h. (What do you do with ?) Hint: Break up the problem into two parts, first consider the Area A of the cylinder base and then use V=A h.

3.4 Repeated measurements and the mean


When you have a group of n repeated measurements, it makes sense to calculate the average value, or mean value for that group. The mean X is simply the sum of all n values, divided by the number of values in the group.

1 X = -- X i n
i=1

(EQ 5)

3.4.1 Calculate the mean for the following group of values: 22.1 cm, 1.5 m, 307.3 mm (Hint: all data which is mathematically combined must have the same units. As concerning the significant figures, the rules for addition apply. Keep in mind that a constant is a number considered to be WITHOUT error.)

The mean, too, must always be recorded with its associated error. But what is the error of the mean?

3.5 The error in the mean: the standard deviation


What would be a logical way to discuss the error associated with the mean of a group of values? Each value was taken with the same instrument, perhaps just the uncertainty in the measurements would suffice. Alas, no. Typically the mean deviates from the data by a value not simply related to the uncertainty. A better way would be to determine the average deviation of each individual data point from the calculated mean.This is known as the standard deviation n1:

EXP I

n 1 =

2 1 ----------- ( X i X ) n1 i=1

(EQ 6)

Note that although there are n data points, n1 is used when taking the average. This makes n1 a so-called unbiased estimator. 3.5.1 Calculate the mean and standard deviation for the following data: .019 gm +/-.005 gm, .019 gm +/-.005 gm, .021 gm +/-.005 gm.

3.5.2 Is the standard deviation larger or smaller than the uncertainty in the values? Does this make sense?

3.6 Experimental error


Errors associated with experiments may be summarized by three types: Systematic Errors - These include prejudice on the part of the observer, inherent defects in equipment, neglect of effects such as temperature, pressure, humidity, etc. These errors are characterized by their tendency to be in only one direction. For example, if a meter stick is slightly worn at one end and measurements are taken from this end, then a constant error will occur in all measurements. Unknown systematic errors can only be discovered by comparing measurements of the same physical quantity which are obtained by several methods. Personal Errors - Error due to blunders, such as mistakes in arithmetic, in recording data, or in reading measurements. These can be minimized by repeating measurements and keeping redundant record which are subject to crosschecks. Random Errors - A series of measurements in which the systematic errors have been minimized will still contain variation due to causes that lie beyond the control of the observer. These are characterized by discrepancies in several measurements of a quantity under apparently identical conditions. It is assumed that they are due to the combined effect of a great number of independent causes, each of which is equally likely to produce a positive or negative effect. Random errors are statistical in nature and therefore can be analyzed by statistical methods.

EXP I

3.6.1 From your previous experience(s) in science labs, give an example of each type of experimental error.

3.7 Expressing the experimental results


The number of significant figures for the standard deviation you calculated is the same as that for the data points. But when you express the experimental error as

X n 1 , it would make more sense to keep only two (or maybe three)

significant figures for the standard deviation or anything else used as an experimental error. For example if the mean and standard deviation are 2464.561 and 3.278695 (seven significant figures each), then the result should be expressed as

2464.6 3.3 , or, alternatively, 2464.6 0.13% , or, in scientific notation, 2.4646 10 0.13% . Note

that the number of digits kept after the decimal point for the experimental result should be the same as for the error.

4.0 Graphing data


Graphical analysis of a data set is an extremely useful tool to the scientist. Often the general shape of the curve described by the data gives clues to whether you did the experiment properly and if your data supports the theory. Properties of lines and curves can give you information about the theory from a set of graphed data as well. Thus it's important we're all speaking the same language with respect to graphs.

4.1 Graphical syntax


Before attempting to construct a graph, your data must be organized into a clear and easily readable form. It is usually desirable to put your data into the form of a table, clearly indicating what was measured, the units and comments about experimental uncertainties. If some data seems to clearly disagree with the rest, put a question mark next to it, but do not eliminate it at this time. A graph can present a clear picture of how one physical quantity depends on another. However, to be effective, a graph must conform to accepted rules. In order to display the information clearly, choices must be made concerning the size of the graph, scale for each axis, appropriate use of symbols and words and the type of graph (linear, semi-logarithmic, etc.). The rules presented below are followed by most scientists and engineers. Use graph paper or an acceptable substitute. Data taken in the introductory labs are often accurate to three significant figures. The graph should be made as large as possible in order to retain this accuracy. Scales for the coordinate axes should be chosen so that the data

EXP I

extends over almost all of a full-sized sheet of paper. It may be necessary to suppress the zero so that the data and the resulting curve will cover most of the graph paper. The decimal parts of units should be easily located. This can be done if each small division is made equal to numbers such as 0.1, 0.2, or 0.5. The scale need not be the same on both axes; however, each axis must be labelled by the quantity and units being plotted and the division used. The units of the measured quantities are customarily enclosed in parentheses. If velocity is being measured in m/sec, use v (m/sec). The use of a power of 10 in a scale caption in the form v in m/sec 103or v 103 m/sec should be avoided since it is not clear whether the scale numbers have been or are to be multiplied by 103. To avoid misunderstanding when it is necessary to use a factor like 103, it should be directly associated with the units as v(103 m/sec) or directly associated with the scale numbers. A brief caption may be inserted in a vacant area within the graph to make the graph reasonably self-explanatory. All symbols used in the graph should be explained. Make sure the graph is titled. In plotting a curve, the dependent variable is plotted along the vertical axis and the independent variable is plotted along the horizontal axis. For example, a straight line through the origin with slope m is represented by the equation y = m x. A plot of this line is shown in Figure 1. 4.1.1 Can time ever be properly plotted along the y-axis? Explain.

Graph of y=mx [slope = rise/run = y/x = m] y(x)

Figure 1: Graph of y = mx

EXP I

Graph of y=mx [slope = rise/run = y/x = m] y(x)

Figure 2: Graph with error bars


Technically, actually connecting graphed points with a solid line/curve indicates that you know that the values between your data points lie along that curve, that it is a fact and you are certain. Dashed or dotted lines/curves indicate you believe the theory to hold between the data points but you dont have the data. Best-fit and theoretical lines or curves are drawn through a group of data points and are labelled as exactly that -- best-fit or theoretical, not exact. This may appear a technicality, but it is part of the graphical language science shares and its wise to adhere to it. A best-fit or theoretical smooth curve should be drawn in such a way as to fit the points as closely as possible within the error bars and, in general, as many points will be on one side of the curve as on the other. The extent to which the plotted points coincide with the theoretical curve is a measure of the precision of the results. If the experimental points lie along a straight line, the location of the best fit line can often easily be found by sighting along the points when the eye is placed almost in the plane of the paper (eye-ball fit).

5.0 A Simple Experiment and Analysis


So lets do some experiment and analysis of something we deal with all the time in our everyday livesgravity. Gravity is an acceleration which is due to a forcethe force of gravity. Seems intuitive that accelerations are always caused by forces being applied. The physical quantity acceleration is defined with respect to position and time. Imagine some object at some position X from an origin of your choice. It moves with a certain rate in a certain amount of time and you are asked How far did it move? In Jr. high you were probably taught that the answer was, distance = speed x time. This is still true. Speed is the magnitude of velocity, which, in addition to the magnitude, has a direction. In one dimension velocity is just the speed and a plus or minus sign in front of it. For objects moving to the right, the velocity is taken to be positive and vice versa. Likewise the change in distance or position is taken positive if toward the right direction. The velocity

is equal to the change in distance X over the change in time t : = X . -----t


(EQ 7)

The velocity tells you how fast an object is changing position,

X . One can also look at how fast the velocity itself

EXP I

10

is changing,

. The change in velocity is known as the acceleration a of an object:

a = . ------t
Notice that both Equations 7 and 8 are in the form graphs.) 5.0.1 From Equation 7 what quantities would you measure to determine the velocity of an object?

(EQ 8)

m = y , where m is the slope of the line. (See earlier examples of ----x

5.0.2 Will a graph help? If so, explain carefully how.

Clearly this is where the experiment is heading, but first we need to learn to use our experimental equipment, the motion sensor.

6.0 Motion Sensor


The motion sensor is one of a number of different types of sensors to detect position. It sends out an ultrasonic (of frequency greater than you can hear) sound pulse at regular, preset time intervals. The sound bounces off an object and returns to the detector. Using

x = vt , the motion sensor then carefully determines the distance to an object, here x v 345 m/s the speed of sound in air at

and t being the round-trip distance and time, respectively, and

T = 73 F .

11

EXP I

6.1 Setup
1. If the black PASCO interface (next to the computer) is not already turned on, turn it on. If the interface is not the USB type (but the SCSI type), the computer must be restarted after it is turned on. Connect the motion sensor to the PASCO interface. The yellow plug goes into Channel 1while the black plug goes in Channel 2. 2. Start the DataStudio software. 3. Welcome to DataStudio window should pop up. Choose Create Experiment. (If this window doesnt appear, click Setup.) The Experiment Setup window will show you the PASCO interface. 4. Click on Channel 1 on the interface and add a Motion Sensor. An icon of the motion sensor should appear with the plugs already connected to the correct channels. Double-check that the yellow plug is connected to Channel 1 and the black plug is connected to Channel 2. Note that if you accidentally select the wrong sensor and wish to delete it from the screen, simply click on that sensors icon and then press the Delete key on the keyboard. Finally in the Experiment Setup window, change the Sample Rate to 20 Hz. This simply means that the computer will collect 20 data points, equally spaced in time, every second.

6.2 Taking data


Generate a Graph display as follows. In the left column in DataStudio, you will see Position, Velocity, and Acceleration variables under Data. Click on Position and drag it onto Graph, which is under Displays in the same column. This will generate a Position vs. Time graph. Then click on Velocity and drag it onto Graph 1 (but not Graph) under Displays. Finally click on Acceleration and drag it onto Graph 1 (but not Graph) under Displays. This should generate the Position, Velocity, and Acceleration vs. Time graphs with their time axes aligned, i.e., Graph 1 should consist of three graphs that are vertically stacked. After you finish generating your graph, maximize it to full screen. If you have done everything correctly, your graph should look like the one in Figure 3 below.

EXP I

12

Figure 3: DataStudio x, v, a vs. t graph


To take data simply click the Start button. Click the Stop button to stop taking data. In order to erase unwanted or junk data, choose Delete ALL Data Runs from the Experiment menu. You should delete data runs as necessary so that the program doesnt crash or you dont get confused by many different data runs. 6.2.1 Take a data set while one partner is slowly moving back and forth in front of the motion sensor. Sketch the graph of position vs. time.

13

EXP I

7.0 Exploring position, velocity, and acceleration


In this section you will make qualitative and quantitative studies of position, velocity, and acceleration using the motion sensor. 7.0.1 With one partner holding the motion sensor still, have the other partner move in front of it such that the data you take for real mimics the following graphs shapes. (Dont worry about numerical values here...) For each graph, explain how you had to move to replicate the graph shape (NOTE: x=position, v=velocity). t Graph B Graph A

Graph C

Graph D

EXP I

14

7.1 Abruptly changing velocity


7.1.1 Given the x vs. t plot in Figure 4 below, describe how the object is moving, i.e., give the speed and direction, in each lettered region. You could alternatively give the velocity with the magnitude indicating the speed and the sign indicating the direction (plus to the right, minus to the left) in each region. Dont forget the units.

7.1.2 The motion sensor records increasing distances from the sensor itself as positive. Therefore, if the motion sensor is facing toward the right, the increasing right direction will record as positive data points on the computer, with the sensor itself acting as the origin ( x = 0 ) of your position scale. Now use your laboratory partner as an object and quantitatively produce Figure 4 on the computer by collecting the data with the motion sensor. The motion sensor should sit steady and your partner should move. Use a meterstick and stopwatch if necessary.

7.1.3 What do you expect the acceleration to be when you are within one of the lettered regions in the graph?

7.1.4 What is happening to the acceleration when you switch from one region to the other? How do you explain this? Is it physical?

15

EXP I

3.0 C D E F

x (m)

2.0

1.0

0.0

10 t (s)

11

12

Figure 4: x(t) plot for an object with abruptly changing velocity.

7.2 Slowly changing velocity


7.2.1 Given the x vs. t plot in Figure 5 below, describe how the object is moving, i.e., give the speed and direction.

7.2.2 Now use your laboratory partner as an object and quantitatively produce Figure 5 on the computer, by collecting the data with the motion sensor. The motion sensor should sit steady and your partner should move. Use a meterstick and stopwatch if necessary.

EXP I

16

7.2.3 How is the velocity graph behaving now? Explain how your velocity graph relates to the position graph on the computer.

7.2.4 How is the acceleration graph different this time?

7.2.5 Do you see the connection between the velocity and acceleration graphs? Explain how the two graphs are related to each other.

17

EXP I

3.0 C D E F

x (m)

2.0

1.0

0.0

10 t (s)

11

12

Figure 5: x(t) plot for an object with slowly changing velocity.


We will continue this experiment in next weeks laboratory session; so refresh your memory before you come to the laboratory.

EXP I

18

Name: Partners name: Date and time:

PHYS135AL

EXPERIMENT II DATA ANALYSIS: APPLICATION

1.0 Objective
You will apply the methods you learned last week and determine a very important physical quantity, the acceleration of gravity near the surface of Earth.

2.0 Linear data


You are probably already familiar with linear graph paper. Quantities along an axis are directly proportional to their linear distance from the origin. The coordinates of the origin are often, but not necessarily, zero. As an example of data that can be linearized via graphing techniques, consider an experiment in which we measure the acceleration due to gravity on Earth. The force of gravity is caused by the mass of the Earth, and is directed straight down toward its very center. The acceleration is in this same direction. So objects in free fall, falling toward the Earth unencumbered, are accelerating due to this force alone. For objects near the Earths surface, this value for the acceleration g is constant. In MKS units g = 9.81 m/s2. In this experiment, using the motion sensor to determine the distance of a moving object accelerating under the force of gravity, you will find a value for g and compare it to the accepted value. Figure 1 shows the experimental setup. A rod is clamped to the side edge of the table, which is used to set the dynamics track at an inclined angle. The 0.0 cm end of the dynamics track should be raised about 2530 cm with respect to the opposite end. Too much angle will result in the cart going wild and derailing. Too little angle will result in insufficient acceleration. The bumpers at both ends should have their magnets facing inside. The clamp holding the rod should be on the opposite side of where the knob of the upper-end bumper is. Both bumper knobs should be securely tightened. The motion sensor is attached to the upper end. A small rod through the motion sensor further secures it at the bottom of the track with a duct tape. The yellow and black plugs of the motion sensor should be connected to Channels 1 and 2 of the interface, respectively.

3.0 Procedure
Apparatus Dynamics track, collision cart, Motion sensor II (CI-6742) on a short rod attached to the bottom of the track with a piece of duct tape, table clamp and rod to raise the dynamics track, magnetic bumper with the magnetic side facing in at the bottom end of the inclined track, right-angle triangle, interface (ScienceWorkshop 750 USB), and computer with DataStudio. The cart has nearly frictionless wheel bearings. Before you start take the cart in your hands, turn it around, and center the wheel axes if necessary by carefully pushing the wheels sideways with your fingers. Adjust the angle of the motion sensor to 0 degrees, using the large knob on its side. This adjustment is important;

EXP II

Figure 1: Dynamics track set at an inclined angle


otherwise, the experiment may not work very well. The beam switch should be set to narrow. There should be one magnetic bumper at the bottom end of the track with the magnet side facing in. No magnetic bumper is needed at the top end of the track. Before you start read these cautions carefully: CAUTION: Treat the dynamics tracks and carts with care. Be careful not to hit or scratch the tracks. Do not remove any parts from the dynamics tracks. Do not drop the carts. Have your partner catch the cart at the end of the run. CAUTION: Never hit or drop the motion sensors. Never remove the magnetic bumper in front of the motion sensor. The dynamics cart hitting the motion sensor might damage it. CAUTION: Always have your partner (or yourself) catch the cart before it hits the bumper at the end of the track. The cart will probably travel too fast to be stopped by the bumper alone. Also make sure that the carts do not somehow end up hitting the computers or computer accessories. CAUTION: The carts contain powerful neodymium magnets. Never bring them close to the computers, hard drives, or floppies.

EXP II

Also the dynamics tracks should be periodically wiped with a soft cloth to keep them clean so that the carts ride on the tracks smoothly and with minimum friction. Now start DataStudio and click Create New Experiment. Under sensors double-click the motion sensor, which will assign it to Channels 1 and 2. Then double-click the motion sensor icon and, under the Motion Sensor tab, change the trigger rate to 20 Hz. 3.0.1 20 Hz equals how many cycles per second?

3.0.2 What then is the time interval between pulses for a frequency of 20 Hz?

Now in DataStudio drag Position under Data in the left column to the Graph under Displays to generate a position-vs.time graph. Then drag Velocity under Data in the left column onto the newly generated graph, releasing the mouse only after the entire graph is highlighted. This will put a velocity graph under the position graph, with the time axes of the two aligned.

3.1 Qualitative analysis of acceleration


Last week you were asked to replicate the shapes of displacement and velocity graphs by moving in front of the motion sensor. Go back to those graphs and do the following: 3.1.1 For graphs A and B from 7.0.1 in Experiment I, qualitatively describe what the velocity of the object is for the different parts of the graph. Hint: what does EQ 7 in Experiment I tell you?

3.1.2 For graphs C and D from 7.0.1 in Experiment I, qualitatively describe what the acceleration of the object is for the different parts of the graph. Hint: what does EQ 8 in Experiment I tell you?

EXP II

3.1.3 Are any of these four graphs linear? How do you know for sure? Explain.

Now click the Start button in DataStudio. Then, moving the cart on the dynamics track with your hand, try to qualitatively replicate the following two graphs. You will need to use the pan and zoom tools on the graphs to focus in the regions of interest. After you are done, click the Stop button. (Hint: For Graph b let the cart make a gentle bouncing motion off the lower-end bumper.)

a V X

3.1.4 For each graph, if you suceeded, explain carefully how you replicated these graphs. If you failed, explain why you could not do it.

3.1.5 Draw two different graphs (x vs. t and v vs. t), which someone would not be able to replicate.

EXP II

3.1.6 In your own words, explain in general how graphing data can help you demonstrate or describe physics equations.

3.2 Quantitative analysis of acceleration


For this activity set the angle of the inclined between approximately

10 and 5 .

3.2.1 Choose two points as far apart as possible (Why?) along the rule of the dynamics track and record the distance L (measured along the track) between these points. Include the error and unit as well.

3.2.2 Using a meterstick measure the height of these two points from the table and write them down with the error and unit. Then find the difference of height for these points, H, recording it with the error and unit. Is the error for the difference the same as the individual measurement errors?

Here is how you will take data. The minimum distance the motion sensor can measure is about 15 cm. Therefore release the cart from about 20 cm. Do not let your hands or anything else block the view of the motion sensor. Sometimes the motion sensor may miss the target (cart), resulting in very large readings. You should make several runs until you are satisfied that you have good data. You can always delete unwanted runs using the Edit menu. Now click Start and have your partner release the cart. Make sure that your partner catches the cart before it hits the lower-end bumper. You can click Start before the cart is released, but you might want to click Stop before the cart is caught in order to avoid chaotic data points at the end of the graph.

EXP II

3.2.3 Are all your data points good? Which ones can you ignore? Why?

4.0 Determining g
The computer displays the velocity of the cart at different times by simply using successive values of the distance and dividing their difference by the time interval which you calculated in 3.1.2.

Remember the definition of acceleration,

a = V ( means change in here). The expression for a is a linear -----t

relationship. Given good data one expects to see (at least roughly) a line when its graphed properly. What do we remember about lines? Well, y = mx + b, of course, where y is the dependent variable and x is the independent variable. Then m is the slope of this line, and b is the y-axis intercept. 4.0.1 Look carefully at the equation for the velocity, v=v0+a t. Does it fit the y=mx+b pattern? What are y, x, m, and b?

4.0.2 For about 20 evenly spread good data points, set up a table for velocity vs. time. Use the xy tool on the graph to determine the velocity and time at each point.

EXP II

t (s)

v (m/s)

4.0.3 Create a graph of your own design on linear graph paper that will give you a line whose slope is the acceleration of our cart on the dynamics track. Use the rules in Experiment I. Whether you pass or not will be determined by your thoroughness and neatness.

EXP II

4.0.4 Carefully make a best-fit guess and draw this line on your graph.

The purpose of a best-fit or worst-fit line through a set of data points is to determine your expectations and just how good or bad your values are compared to these expectations. A worst-fit line is the line which deviates from your bestfit by as large a slope as possible but still fits your data within reason. 4.0.5 Carefully make a worst-fit guess for the steepest line and the least-steep line which may fit your data. Draw them on the graph and label them.

To obtain the slope m = y/x, choose any two points on the best-fit line. Then divide the change in y, y = (y2 y1), by the corresponding change in x, x = (x2 x1). It is important to remember to include the scale factors and units for the quantities y and x in calculating the slopes, otherwise the value of m determined will be incorrect. It should be emphasized that in choosing the two positions on the line for calculating the slope, the experimental points

EXP II

should not be used if a best-fit line has been found. After the line has been drawn, ignore the points and find the slope using two arbitrary positions on the line. Also make sure that these two points are as far apart from each other as possible (Why?).

4.0.6 What is the slope of the best-fit line?

4.0.7 What are the units of the slope? Is this slope a physical quantity we should recognize? If so, what is it?

Once we have determined the slope of the line, the acceleration due to gravity can quickly be determined by the following trigonometry:

g len

th L

height H

Remember your trigfor right triangles the definition of sin O = opposite divided by hypotenuse.

4.0.8 Determine sin = H/L. From here use EQ 1 to determine the acceleration of gravity g from the acceleration a of the cart on the dynamics track:

a = g sin

(EQ 1) (EQ 2) (EQ 3)

4.0.9 What is the error in g? (Use best-fit and two worst-fit lines to determine the error.)

4.0.10 What then are the possible values of g according to your experiment?

EXP II

4.0.11 Which one of the possible values is the best value? Explain.

5.0 Least-squares fit


In the previous example, it was easy to draw a straight line through the data points by eye and obtain a satisfactory fit to the data. However, there are experiments where the data will show a greater spread, and several straight lines might be a best fit. In such cases, there is an analytical procedure based on the statistical principle of least squares that will yield the best fit to the data. This is the method of least squares. Consider a set of N data points that represent measurements of the dependent variable y as a function of x, y = f(x). (Were going to ignore the mathematical technicalities here. Youre certainly welcome to look them up on your own...) Each data point is given by the pair, (xi, yi). The least-squares fit to a straight line is a method for determining the most probable values for the coefficients m and b, considering the relationship y = f(x) to be linear and of the form yi=mxi+b. The best values of the coefficients m and b are obtained by a tedious minimization technique which gives:
N N N N ( x y ) y x i i i i i=1 i=1 i=1 m = ----------------------------------------------------------------------2 N N 2 N x i x i i = 1 i=1 N 2 xi N N N

yi xi ( xi yi )
i=1 i=1 b = i=1 i=1 ---------------------------------------------------------------------------2 N N 2 N x i x i i = 1 i=1

Theyre ugly, and a bit of patience is required to calculate them (it helps to note the denominators are the same), but when best-fit via eyeball isnt really easy, this is by far the best method of the two. 5.0.1 For your data, calculate m and b using three data points of your choice and the method of least squares. What does m represent physically?

5.0.2 Determine the acceleration of gravity g from this slope and compare the least-squares-fit value for g to the best-fit-eyeballed guess g.

EXP II

10

5.0.3 For the dynamics-track experiment, list one possible source of systematic error that may affect your calculations. Based on your choice of systematic error, would your experimental value of g be larger or smaller than without the error?

5.0.4 Suppose we increase . Would the experimental values of the acceleration a and gravitational acceleration g change?

6.0 Studying friction


In introductory physics the frictional force is usually said to be proportional to the normal force. This is not necessarily true for a cart on wheels. See the free-body diagram in Figure 2.

Figure 2: Free-body diagram for the inclined plane.


N = mg cos Ff mg sin

mg cos

mg

If the frictional force

F f is proportional to the normal force N , then this means that it is proportional to cos . But sin .

in reality, for a wheeled cart, it could be in part proportional to the parallel component of the gravitational force. In this case you could have a frictional force proportional to

6.0.1 In this activity you will vary the angle to study the variation of the frictional force. For eight angles between

11

EXP II

approximately 10 and 2 , do DataStudio runs and fill out Table 1. Before each run measure and record the height difference H and the length L along the inclined plane to millimeter accuracy. You can use a constant length L for each run, such as between the 10 cm and the 220 cm mark on the rule, and measure the height difference for these marks. Use the right-angle triangle provided to you to help with the measurement. In order to get the acceleration a, this time make a Linear Fit to your velocity vs. time curve, using the Fit icon on the graph. The slope will be the acceleration a along the inclined.

Table 1: a vs. . L H sin = H/L a

When done filling in the table, choose New Empty Data Table from the Experiment menu. For the x and y values, enter the values of sin and a, respectively. After the last entry, make sure to press enter. Then drag Editable Data in the left column under Data onto Graph under Displays. This will generate an a vs. sin graph. 6.0.2 Make a Linear Fit to the a vs. sin graph, using the Fit icon. Determine the gravitational acceleration g from the fit, and record it with the experimental error (standard deviation) shown in the fit.

EXP II

12

6.0.3 Record both the slope and intercept, along with their experimental errors shown, from the fit. Using the known value of g = 9.81 m/s2, determine the parts of the frictional force that are proportional to sin and cos . Note that cos 1 for the small angles you studied.

6.0.4 Is the part proportional to cos greater than the error? Could you then say the frictional force proportional to the normal force is nonzero?

6.0.5 What did we assume in the estimation of the frictional force in 6.0.3? Then what errors could contribute to this estimation?

13

EXP II

EXP II

14

Name: Partners name: Date and time:

PHYS135AL

EXPERIMENT III TWO-DIMENSIONAL MOTION

1.0 Objective
You will use a self-made video to study and make physical measurements of two-dimensional motion, and to gain an intuitive feel for the relationship between data, graphs, and the real world.

2.0 Background
Before you begin this lab, you should have a general idea of how to describe the motion of objects using mathematical terms. If you truly understand this process, you will be able to predict the motion of certain objects if you are told in advance about a few parameters. It is these kinds of problems which you have seen as homework assignments and will continue to see as exam questions. A general review of kinematics follows: Kinematics is the part of science which allows us to describe how objects move. Use of kinematics will lead to precise descriptions of an objects motion. This description may include values of position, velocity, or acceleration at an instant in time or averages of those values over a time interval. The specifics of calculating these values for any given situation depends on the use of certain equations. These equations will contain variables for position (xo,x), velocity (vo, v), acceleration (ao, a) and time (to, t). The subscripted variables denote the value at time t=0. For the purposes of this discussion, we will also limit ourselves to a constant acceleration (ao = a for all time). This leaves us with 6 variables which can be combined together into useful equations. The equations (known as kinematic equations) are derived in your textbook and presented here.

v = v o + at 1 2 x = x o + v o t + -- at 2 v = v o + 2a ( x x o )
2 2

(EQ 1) (EQ 2)

(EQ 3)

Remember that these Kinematic Equations were derived for constant acceleration and are only valid in that instance. An important simplification of these equations occurs when the acceleration equals zero. Now that you remember what the kinematic equations are, lets use them. You will notice that the equations are written for only one-dimensional motion (only x and no y). This is an important feature of kinematics which allows us to decouple dimensions. In english, this means that we treat each dimension separately and use the equations twice, once for each dimension.

EXP III

2.1 One Dimensional Example


First a simple example in one dimension. Suppose a chef drops an egg from a height of one meter. The first question to ask is What do I know? The following values are known simply from the statement of the problem: yo=1 meter, y= 0 meters, vo=0 meters/second, a = g = 9.81 meters/second2. Notice that the choice for y and yo are arbitrary with the important value being (y-yo) = -1 meter. The only unknowns are v (the final velocity of the egg after falling 1 meter) and t (the time that it take to fall 1 meter). These values can be calculated using EQ 1, EQ 2, and EQ 3. The other important function of these equations is to help us interpret graphs. For this example, it may be useful to make a graph of position vs. time, velocity vs. time, and acceleration vs. time. 2.1.1 Make a sketch of what you think the data will look like for the chef/egg example. At this stage, make your best guess and then later you can see how well you did.

2.2 Two Dimensional Example


Now suppose the chef (angry at dropping an egg) impulsively throws another egg across the kitchen. Suddenly, the problem becomes two dimensional. A sketch of this scenario is shown in Figure 1.

Figure 1: Chef throwing egg

EXP III

The obvious questions to ask are: 1. Will the egg actually reach the assistant chef? and 2. How much time does the assistant have to move out of the way? These questions can easily be answered using our knowledge of kinematics. In this case, the problem is broken into two independent problems. Let us again ask What do I know?, this time for each dimension. You will notice that we do not know very much (in fact, we only know ax=0 and ay = -g =- 9.81 m/s2). So lets make some assumptions to make the problem more interesting. In the real world, you would have to make some measurements to find more information. In this case, let us just say that the chef and his assistant are standing 10 meters apart and that the chef launches the egg with an initial velocity of 10 meters/second at an angle of 30 degrees above horizontal (intending to throw the egg over the head of his assistant of course). Now, with these additions to the problem, you should have enough information to complete the problem and answer our two initial questions. Does the egg hit the assistant? and How much time does the assistant have to move? Start with the horizontal component since we know the most information. Using the equation for the position as a function of time (EQ 2) in the x-direction (horizontal direction), we find 10 meters= 10 meters/second * cos(30o) * t, or t=1.15 s. This is the time it takes the egg to travel 10 horizontal meters. Now try to answer the next question. 2.2.1 Apply the equation for the position as a function of time, but this time in the y-direction (vertical direction). Determine and plug in the y-component of the initial velocity and acceleration in the y-direction to find the relative height at the instant the egg reaches the horizontal position of the assistant (t=1.15 s). So does the egg hit the assistant? If it does, at what part of his body?

2.2.2 Once again, sketch graphs of position vs. time, velocity vs. time, and acceleration vs. time for this example. Include one such graph for each dimension (6 total). Remember at this stage the important idea is that you make a good guess about the shape of the graph.

EXP III

2.3 Graphs
Now that you have sketched a series of graphs, lets quickly review why these will be useful. For any graph, one can calculate the slope and the area under the curve.This is especially useful if it is easy to do and the result is a physically meaningful value. Figure 2 shows two graphs. The question to ask is, What could possibly be useful in these graphs?.

t Figure 2: Sample Graphs

First look at Graph A. Immediately you could calculate the slope and y-intercept for this line. Physically, the yintercept is the position y at time t=0. What is the slope? Remember slope = rise/run so slope = y/t which is just the average velocity. Is the area under the curve useful? That area would be some quantity called meter*seconds. In this case, the area under the curve is probably not useful. Now look at Graph B. Again you can calculate the slope, y-intercept, and area under the curve. In this case, the slope is v/t = average acceleration. The y-intercept is the initial velocity at time t=0. The area under the curve is some quantity with units of (meters/second) x second (which is just meters). So it turns out that for a v vs. t graph, the area under the curve is the net displacement (a useful quantity). Now that you have reviewed kinematics and graphing, you are ready to proceed with actual measurements. For your convenience, here is a summary: The kinematic equations (EQ 1, EQ 2, and EQ 3) are valid for one dimensional motion with constant acceleration. Two-dimensional motion can almost always be broken into two one-dimensional problems. Graphing is a useful analysis tool.

EXP III

3.0 Procedure
The procedure this week will include shooting a video, capturing the video, editing the video, analyzing the motion of an object in the video, and studying the accompanying graphs of the motion. You will use a simple motion scenario this week (bouncing racquetball) to familiarize yourself with the operation of the camera and software as well as to understand the interpretation of the associated graphs which the software will generate. A diagram of the setup is shown in Figure 3.

Chefs Hat not required

Racquetball

Meterstick Figure 3: Experimental setup

EXP III

3.1 Shooting the video


Your first task will be to film some motion. The two most important considerations to account for are parallax and camera motion. Parallax: The motion you see on a video screen is two dimensional. Therefore, to get accurate measurements you must insure that all motion you shoot is in two dimensions. [see Figure 4]

Wrong
Plane of motion

Right
Plane of Motion

Camera

Camera

Figure 4: Parallax
The easiest way to insure this is to stand far from the object of motion. Only if necessary, use the ZOOM button to close in on the object. Camera Motion: It is important for the later analysis of your video that the camera does not move during recording. Setting the camera on a table or chair is often a useful way to keep the camera still.

The video camera you will use is a standard MiniDV digital camcorder. The following procedures outline the use of the camera: 1. The battery should be attached to the back of the camera at all times. Make sure that the P/ EASY switch is in the P position. The media switch should be in the tape position. Press the little gray button and while it is pressed, turn the camera power dial to CAMERA. Open the lens cover by sliding the switch near the front. Occasionally, if the camera is left unused for a couple of minutes, the camera will shut itself off to save power. In this case, turn the camera power switch to OFF and back to CAMERA. 2. Open the LCD viewfinder and press FUNC. The following settings should already be set and remembered by the camera (it has a backup battery) but verify them just in case, pressing FUNC first and then using the joystick to scroll down to the MENU and pressing it to enter the menu: CAMERA SETUP SHUTTER: 500, A. SL SHUTTER: OFF, D. ZOOM:OFF, IMG STAB: OFF, WIDESCREEN: OFF VCR SETUP REC MODE: SP, TV TYPE: 4:3 DISPLAY SETUP TV SCREEN: ON, MARKERS: GRID (GRY). After done, while pressing the little gray button, slide the control switch to PLAY and make sure TV SCREEN is set to ON under DISPLAY SETUP in the MENU under FUNC. (Note: The shutter speed sets the time period for which
the shutter is opened per frame of a shot. High shutter speeds prevent motion in the video from appearing blurry. However, high shutter speeds also require more light for illumination. Generally a shutter speed of 1/500 seconds

EXP III

is adequate for indoor recording of motion. If you go outdoors, you may be able to use higher values to get crisper motion.)

3. While pressing the little gray button, slide the control switch to CAMERA again. In order to start recording, press the RECORD button (the big white button located by your thumb on the back of the camera). To stop, press this button again. 4. A meter stick (or some other object of known length) should be placed in the area being filmed. The full length of the object should be in view during the entire sequence filmed. The ball should also be in view the entire time. The widest available (W) zoom usually works best. NOTE: Try to record the motion in the center of the frame. This will be useful later when digitizing the video. 5. A sequence of only a very few seconds of motion will be required for analysis. You may want to record 2 or 3 separate sequences of motion, so that you can choose the best one when you are ready to capture the video on the computer. 6. Notice that these cameras have tape counters showing the remaining amount of tape in whole minutes. This will help you find your segment of recorded video when you are ready to capture. Do not simply assume that your video segment is at the beginning of the tape. There is also a precise counter for individual video segments recorded on tape. 7. Once you are done, turn off the camera by sliding the control switch to OFF while pressing the little gray button. Then, close the lens cover by sliding the switch in the front.
[Technical note on cameras and video formats for the interested reader: All consumer analog and digital camcorders, including the highdefinition models marketed as progressive scan, eventually convert the video to an interlaced format; hence, its crucial for our real-time motionanalysis purposes that the capturing process can automatically deinterlace the video. This is because with interlaced frames, you can have double images of the ball in each frame, corresponding to the upper and lower fields, separated by one vertical pixel shot at separate, equally spaced times and interlaced together as a single frame. In addition this camera can also record with progressive scan (noninterlaced) on SDHC card (not installed) but it does so at half the frame rate (15 fps) and at a 320x240 resolution, with degraded video quality as well. Another problem for real-time motion analysis is MPEG and MPEG-based compression schemes such as H.264 and AVCHD, which compress multiple frames in time together. Separating the frames precisely might not be possible and this may ruin real-time data for the motion. The only formats which dont have MPEG-like compression are standard-definition MiniDV and Motion JPEG. The former is the format our cameras use and the latter is the format used to shoot low-quality, slow-shutter-speed standard- or high-definition videos in still cameras.]

3.1.1 Capturing and editing the video Some of you are probably already familiar with capturing videos on your computer. The process involves transfering the video to a computer file. Here is how: 1. Connect the camera to its DC power supply (adapter). This will save the battery for future use by other students. 2. Use the supplied IEEE 1394a cable to connect your camera to one of the IEE1394a ports on the back of your computer: Gently lift the large gray cover from its notch for the AV, USB, and DV sockets and, after aligning it in the correct orientation, attach the smaller 4-pin connector to the DV socket (rightmost socket) on the camera and the larger 6-pin connector to one of the IEE1394a ports (labeled 1, 2, and 3) on the back of the computer.

EXP III

3. After pressing the little gray button, turn the control dial on your camera to the PLAY mode. Wait for a few seconds for the computer to recognize the camera. You can cancel any pop-up window asking for an action if it appears. 4. The only software that you will need is VideoPoint Capture 2.2.0 (or higher). Start VideoPoint Capture 2.2.0 (not VP Capture, which is the old Version 1, or not VideoPoint or VideoPoint Physics Fundamentals, which are analysis software). When the program starts, make sure that the following Preferences are checked under the Edit menu: De-Interlace Video, 320 (Sony) (even though we are using Canon, still check 320), and Allow 640x480 Movie Capture. Also make sure that the selected Camera is Microsoft DV Camera and VCR.
(Note: In very rare cases, with multiple video devices attached to the computer, if Microsoft DV Camera and VCR doesnt stick after its set under Preferences but another video device sticks, manifested by seeing the preview by another video device when you attempt to Capture, click on Capture if not already clicked and set the video device to Microsoft DV Camera and VCR by clicking the Settings button under the Record button, and then exit and restart the software.)

5. Click on Capture. 6. Play, rewind, pause, or stop your video using the small joystick on the LCD display of the camera. Note that you can also rewind and forward in the play mode. When your video is played, it should also show on the computer window. 7. When you find the correct video segment, click on Record just before the action starts. Click on Stop when the desired action ends. 8. On the next screen, choose a frame count under Adjust Frame Count. We recommend that you choose All. You can also choose Double (frames doubled by combining the two interlaced fields together) but this could likely result in more noise when you analyze your data. 9. Now edit your video by using the left and right perpendicular triangles under the preview window . Slide the left triangle to the point just after a single parabolic trajectory starts. Slide the (a) Video recorded on tape. (b) Edited video. right triangle to the point just before the same single parabolic trajectory ends. You Figure 0-1: Editing the 2D video for the should make sure that there are no bounces trajectory of interest (contacts with the floor) within your trimmed video segment and the trajectory is a nice parabola such as in Figure 0-1. 10. Click Confirm Edit after youre satisfied with your trimming of the video. Then go to the next screen by clicking on the next-step button . You can skip this screen by clicking on again and go to the screen where there is the Save Movie button. (Note that the recommended compression is
Sorenson Video 3 with Best frames per second. Some compression codecs may not work with VideoPoint motion-analysis software or produce

Click this button and then save your video in the My Videos folder under My Documents. You are done with capturing and editing your video.
incorrect data; so, these default settings should be used.)

EXP III

3.2 Analyzing your video


Now that you have a segment of motion in digital form, you are ready to analyze your hard work. Choose VideoPoint under startAll Programs to start the analysis software. Click the close box in the top right corner

of the credits window to begin. Choose the Open Movie option. Find your video segment and open it. You will be prompted for the number of items whose motion you will be tracking. Since your video segment consists of a single bouncing ball, enter 1 and select OK. Three windows will appear: a MOVIE window (with your video segment displayed), a COORDINATE SYSTEMS window and a TABLE window. Each of these windows is interrelated and as you make changes in one window, those changes will be reflected in the others. Make sure to maximize (make full-size) both the VideoPoint main window and the movie-preview window inside it to obtain the best resolution, hence best accuracy.

Figure 5: TABLE and COORDINATE SYSTEM windows


Since the size of the objects displayed on your screen are obviously much smaller than the corresponding sizes in the real world, the first order of business will be to set the scale for the video. To do this, in the menu select Movie : Scale Movie. In the pop up window, enter the length of the scale object which you placed in your movie. If this object was a meter stick, then the length will be 1.00 m. The Origin should be <Origin 1> and the Scale type should be Fixed (i.e. the scale will not change during the movie). Click Continue.

Use the cursor

to select one end of the object in the movie frame. Then select the other end of the object.

The Coordinate Systems window should now display a scale factor relating pixels to meters. If you missed the end of the object when clicking, you can adjust the scale by clicking on one of the scale markers and dragging it to a new, corrected location.

EXP III

Note that to delete any line of information in TABLE or COORDINATE SYSTEMS, highlight the entire row and go to the menu and select Edit : Delete Selected. Double-check the time information for your video. The Table window will have 3 columns of data (time, x-pos, y-pos). Each row of data corresponds to a single frame of video data. The time data indicates the time elapsed between each frame. The data entered in this column should agree with the capture rate which you set in the digitalization procedure. If it does not agree, you must manually set the time data by selecting in the menu Movie : Select Frame Rate and choosing the appropriate frame rate. To do any motion analysis, you must define a reference frame. This usually includes defining an axis and a zero point (origin). In the case of the video point software, you will need to determine whether you need a fixed origin or an origin that moves from frame to frame. The way to decide which type of origin you will need is to play your movie segment frame by frame. If an object exists which does not move in the movie, then you can use this object as a fixed origin. If all objects move a little bit from frame to frame (e.g. you jiggled the camera while shooting the video) then you will need to use a floating origin. NOTE: It is much easier to use a Fixed Origin. Therefore, it is to your advantage to place the camera on a solid object while recording. Fixed Origin: Do nothing. The yellow axes displayed on the movie belong to Origin 1. If you double click the origin, a window will pop-up which shows the parameters of the origin. You can choose to rotate the origin if you like, or use polar coordinates. However, simply leaving the defaults will most likely be the easiest. Floating Origin: Select Create:Origin. Name the point Origin2, set the coordinate type to cartesian, data type to Frame-by-Frame and make sure the Is An Origin box is checked. Click OK. Rewind your video to the beginning (slide the play bar to the far left) and use your cursor to click on the object. Once you have clicked, the movie will automatically advance to the next frame. Click the origin object again and continue for each frame until the movie is finished. To review your choices, play through the video one frame at a time. If you need to adjust the location of a point, do so by dragging the marker to the correct location. IMPORTANT: Click on the object (ball) only during a single bounce (within a single parabolic trajectory), after the ball hits the floor and before the ball hits the floor again. You are now ready to tell the software the location of the object of interest in each frame. You can use the default point (Point S1). Rewind your video to the beginning (slide the play bar to the far left) and use your cursor to click on the object which you are observing. Once you have clicked, the movie will automatically advance to the next frame. Click the object again and continue for each frame until the movie is finished. To review your choices, play through the video one frame at a time. If you need to adjust the location of a point, do so by dragging the marker to the correct location. You can also delete unwanted data points at the beginning and end of the bounce (do so definitely if there are any points not inside a single bounce) after you are done.

EXP III

10

Figure 6: VideoPoint toolbar commands.

3.3 Looking at the graphs


You have digitized data and have marked the location of an object in each frame of a movie (did you notice that the table now has x-pos and y-pos data as well as time data?) Videopoint will now allow you to view a variety of graphs for this data. To begin, from the menu select View : New Graph. A "Plot Series" window will pop up in which you need to decide what data elements you wish to plot. Since you did not measure the mass of the ball, none of the variables that include mass will be useful. The remaining variables (time, position, velocity, acceleration) are all valid choices. To begin, choose x-position for the horizontal axis and y-position for the vertical axis and select OK. A single graph appears. Note that after the graph appears, make sure to change the horizontal axis (usually the time but in this case xposition) and vertical axis to get the maximum zoom (scaled to fit the graph) of the data points. The following steps will help you to get familiar with the graphs. Notice the connection between the video segment and the graph. If you play your video (either completely or frame-by-frame) you will see a blue highlight circle on the graph indicating which point is currently on display for the video. Alternatively, you can use your cursor to click on an individual data point on the graph and the video will advance to the corresponding frame. Since this is a y vs. x position graph, you can do a quick check to make sure that you are set up correctly. Advance the video one frame at a time until you are viewing the frame in which the ball is at its highest point. 3.3.1 Is the highlighted point also the highest? From this graph, can you determine the actual maximum height that your ball bounced? What about maximum horizontal distance traveled? (Be careful: Your origin may not be on the floor).

11

EXP III

Notice that if you click the cursor exactly on the axis line, a window appears which allows you to change the minimum and maximum values along the axis as well as the number of grid lines (Ticks) which are displayed. This feature is available for both axes. Notice that if you click exactly on the axis label, a window appears which allows you to change the variable being displayed. Change the horizontal variable to display time (the graph should now be displaying y-position vs. time). 3.3.2 Can you determine the time-of-flight of the ball? Is the ball at its highest point at the halfway-time? Does this make sense? Explain.

Now look at other types of graphs. With the horizontal axis set to display time, select the vertical axis label. Since y-position is currently displayed, select y-velocity and click the ADD button. A second graph appears in the same window. You can expand the window by dragging it bigger with your cursor. 3.3.3 Does this second graph make sense? Advance the video from the beginning until the ball is at its highest point. What is the y-velocity at this point?

The Model feature of the software will allow you to visually calculate equations which correspond to the data that has been plotted. Click the Blue M-tab on the y-velocity vs. time graph. A model window pops up. Since this graph looks like a straight line, you will want to fit a linear equation. Click Apply and a blue line will appear on the graph. This line is a graph of the equation Vy=At + B (remember, your graph is y-velocity vs. time). Adjust the parameters in the A: and B: box so that the blue line is drawn through the data points for the graph. Remember that the A parameter corresponds to the slope of the graph and the B parameter corresponds to the y-intercept. Recalling this should help you guess which direction to change the parameters. 3.3.4 Once you have determined the parameters which best estimate the straight line, make sure you record the

EXP III

12

equation here.

Remember that for the equation Vy=At + B, the slope of this line is where Vy/t is equal to the average acceleration in the y-direction.

V y --------- = A t

3.3.5 Using the information from the model of the previous step, you should be able to determine the value of acceleration in the y-direction, commonly known as g.

Another way to view g is to view the y-acceleration vs. time graph. Again click on the y-axis label of the graph, select y-acceleration, and click ADD. You should now have 3 graphs displayed. This time use the FIT feature of the software (click the pink F on the y-accel vs. time graph) and choose Average as the Type of Fit. This type of fit will average all the y-values for the graph and will present you also with a standard deviation for the calculation. 3.3.6 Does this result agree with the result from the y-velocity vs. time result?

Now that you have investigated three graphs for the y-direction, Select View : New Graph and set up a similar set of three graphs for the x-direction (x-position, x-velocity and x-acceleration vs. time). 3.3.7 Is there any important information that you can find from these graphs? Remember to look at slopes, intercepts and the individual points as you play the movie through frame by frame. Also look at the differences between the graphs in the x-direction and the y-direction. Record all of your findings. The number of observations recorded should be greater than zero.

13

EXP III

The final thing to investigate is the effect of scale on your results. Make the movie window active by clicking on it. Use your cursor to click on one of the ends of your scale marker. Drag this scale marker so the scale is either longer or shorter than it should be. 3.3.8 Watch (and record here) the effect of such a change on the graphs in each direction as well as the data in the table. Take time to play with this feature such that you have at least a qualitative understanding of the relationship between scale and the graphs.

3.4 Analysis
Now that you are familiar with the graphing window, you can look at graphs which will give you information about the motion you captured. In all the graphs that you have to sketch you should be careful about correctly drawing the axis of the graphs and the origin. Start by setting up 2 graphs, one of x-position vs. time and the other of y-position vs. time. 3.4.1 Do either of these graphs track the trajectory of the object? Why or why not?

3.4.2 Set up a graph which does track the trajectory of the object and sketch it.

EXP III

14

3.4.3 Return to your original graphs (x-position vs. t and y-position vs. t). Sketch these graphs.

3.4.4 What does it mean that the x-graph is a straight line and the y-graph is not?

Recalling EXP 1 and EXP 2, the slope of a position vs. time graph provides useful information. That is, the slope is the average velocity. 3.4.5 Find the average velocity for your two graphs.

Now, set up graphs of x-velocity vs. time and y-velocity vs. time. 3.4.6 Sketch these graphs.

15

EXP III

Again, recalling the first two experiments, you know that the area under a velocity vs. time graph is the net displacement. We will now investigate the areas under the graphs in both the x-velocity vs. time and the y-velocity vs. time graph. 3.4.7 Make a rough estimate of the area under both curves.

3.4.8 Do these values make physical sense for the motion you recorded? If not, what is wrong?

3.4.9 Would the choice of the origin affect these values? How and Why? (If you are not sure, change the position of your origin and measure the areas again. This should help you with your answer.)

3.4.10 Measure net displacements in your video window and compare them to the results from your graphing analysis. Do they agree? If not, why not?

Remember, that the slope of a velocity vs. time graph is the average acceleration. Presumably, the x-acceleration should be zero and the y-acceleration should be g=9.81 m/s2, due to gravity.

EXP III

16

3.4.11 Obtain the values of the accelerations from the slopes of your velocity graphs. Do they agree with the expected values?

3.4.12 Now set up x-acceleration vs. time and y-acceleration vs. time graphs. Again sketch these.

3.4.13 How do these graphs relate to the velocity graphs of the previous steps?

3.4.14 Are there regions of your graph which contain accelerations other than expected? Explain what is happening in these regions.

3.4.15 Are these graphs more or less useful as an analysis tool than your previous graphs? Explain why.

17

EXP III

EXP III

18

Name: Partners name: Date and time:

PHYS135AL

EXPERIMENT IV CIRCULAR MOTION

1.0 Objective
To study the characteristics of uniform circular motion.

2.0 Background
Any object moving in a circle of constant radius at a constant speed is said to be in uniform circular motion (UCM). The minute hand on a perfectly calibrated analogue clock is a good example of this phenomenon: the hand never changes length while rotating once every hour. 2.0.1 Give two more examples of uniform circular motion.

Note that in uniform circular motion, the velocity of the object in motion is not constant but is in fact constantly changing direction. A change in velocity is just the definition of acceleration, so an object in UCM undergoes an acceleration:

v2 a = ---r

(EQ 1)

where v is the tangential speed of the object, and r is the radius of the motion. If v is constant as is assumed here, then an object in UCM traveling on a circle of length 2r at f revolutions per second,

v = 2rf .

(EQ 2)

The associated acceleration is properly known as centripetal acceleration, and is always directed toward the center of the motion. Newton's second law states that any object experiencing an acceleration must be acted upon by some force F, specifically:

2 F = ma = mv --------r

(EQ 3)

where a is centripetal acceleration, and m is the mass of the object undergoing the acceleration. Thus an object in UCM is undergoing a force, the centripetal forcethe force keeping the object on its circular path. As an example imagine a ball being twirled about on a string at constant speedthe string keeps the ball on its circular path; hence, the string produces the force required to produce the centripetal acceleration.

EXP IV

2.0.2 What happens if you remove the force, i.e., you cut the string?

2.0.3 What would be the velocity of the ball if the strings length before cut was r and its acceleration a?

2.0.4 The proper name of this velocity is? Does this make sense?

Note that when the string is cut, the ball is no longer in UCM. In UCM the object is moving with constant speed; thus, the time

t it takes to move through equal angles is

constant. Angular speed (also called angular frequency) is a useful quantity in UCM, and is defined as the change of angle swept out per equal time intervals,

= ------ = 2f , t

(EQ 4)

where f is the frequency of the motion in Hertz (Hz). Hertz simply means cycles (revolutions) per second. Therefore frequency f and revolutions per second n are different names for the same quantity. The units of angular velocity are radians per second. The relationship between tangential speed v and the angular speed is:

v = r .
Then the centripetal force, EQ 3, may be rewritten in a number of useful forms, for example:

(EQ 5)

EXP IV

mv 2 F = --------- = m 2 r r
2.0.5 Rewrite EQ 6 in terms of frequency using EQ 4.

(EQ 6)

All of these forms of Newton's second law are equivalent and may be interchanged when convenient.

3.0 Procedure
In this experiment youll use the UCM apparatus shown in Figures 1 and 2 to verify your answers to 2.0.5 and by comparing centripetal force to gravitational force for different values of m and r.

3.1 The Apparatus


The circular-motion apparatus is shown in Figure 1.

Figure 1: Photo and diagram of the UCM apparatus (shown without the motor attachment).
3.1.1 Determine the mass of the revolving heavy mass with the digital balance.

EXP IV

Start with the spring not connected to the heavy mass and secure the cross-arm in the middle of its range. The indicator post should be secured directly below the point on the bottom of the heavy mass. 3.1.2 Measure the radius of the rotation shaft.

The radius of rotation of the heavy mass is the distance from the center of the indicator post to the center of the vertical shaft. 3.1.3 Measure the radius of rotation of the heavy mass.

Attach the spring to the heavy mass as shown in Figure 1. You will use the motor drive to rotate the shaft. First make sure that the positive connector of the motor is connected to the positive output of the power supply and vice versa. Then turn the voltage knob on the power supply all the way down to zero and then turn on the power supply. You can adjust the rotation rate by adjusting the voltage of the power supply. But do not go above 18 V. Adjust the rotation rate to keep the pointed pat at the bottom of the hanging mass passing directly over the indicator post. If you have trouble seeing the alignment, use a piece of white paper as a background to improve the contrast. The mass will be in UCM at this point. 3.1.4 Measure the period of rotation.

Now slowly stop the rotation by slowly decreasing the voltage all the way down to zero. Tie a string to the heavy mass and extend it over the pulley as shown in Figure 2.

EXP IV

Figure 2: Measuring weight needed to pull mass over indicator.


Put the slotted masses on the hanger attached to the string in regular increments until the point on the bottom of the heavy mass is directly over the indicator post. 3.1.5 Measure using the digital balance and record the mass needed (including the hanger) for the point on the bottom of the hanging mass to hang directly over the indicator post.

3.1.6 Calculate the gravitational force the weight of the slotted masses and hanger in 3.1.5 correspond to. (g

= 9.81 m s .)

Repeat the entire experiment for three other positions of the support arm and indicator post, giving data at a total of four different radii.

EXP IV

4.0 Analysis
4.0.1 Compute the force exerted on the system by the slotted masses and hanger for each radius.

4.0.2 From the measurements of the period for each radius, compute the frequency and angular speed of the rotation.

4.0.3 Explain the connection between the two quantities you have tabulated.

EXP IV

4.0.4 Plot the force necessary to suspend the heavy mass directly over the indicator versus angular speed squared times the radius of rotation,

r.

EXP IV

4.0.5 Use your graph to find the value of the revolving mass and compare it to the measured value.

4.1 System Instability


The theoretical analysis of the system we are studying in this experiment is fairly simple. There are two opposing forces, the centrifugal force and spring force. The condition of equilibrium along the radial direction requires that these two forces exactly balance each other. Therefore, we have

m r = k ( r r 0 ) .
Here

(EQ 7)

m is the mass of the rotating object, is the angular speed, r is the radial distance from the rotation axis, k is

the spring constant, and r 0 is the unstretched and uncompressed length of the spring. 4.1.1 Using Equation 7 solve for the radius r at which the system is in radial equilibrium for a given angular speed

. Express your solution in terms of r 0 , m , k , and .

4.1.2 Using the equation you have just derived, find the value of the angular speed

that gives an infinite

radius, expressed in terms of the other quantities. This is the upper bound on the angular speed that you can have for your system. If you increase the angular speed beyond it, the spring will no longer be able to balance the centrifugal forceit will start stretching indefinitely until it breaks, therefore, resulting in an unstable system.

EXP IV

4.1.3 Plot the gravitational force of the slotted masses and the mass hanger as in 4.0.4 but this time only as a function of the radius

r . Then using F = k ( r r 0 ) , determine both k and r 0 from your graph.

EXP IV

4.1.4 Using the measured value for the rotating mass system instability.

m , evaluate the angular speed that corresponds to the

4.1.5 Express the maximum angular speed

that you attained in this experiment as a percentage of the system-

instability angular speed . Comment on how close you were to the system instability in this experiment.

4.2 Real-Life Example: The Earth


Clearly we live on a rapidly-rotating body, the Earth. What are the consequences of this rotation on humans? Consider Figure 3.

Figure 3: Geometry of the Earth sphere Figure 4:

EXP IV

10

4.2.1 Using the wall map outside KAP B-19, find the latitudes of the Equator, Los Angeles and Panama City, Panama.

The radius of the Earth, RE is 6.4 x 103 km. Note that if a person is standing at degrees north latitude (point A), then the distance, r, of that person from the Earths axis of rotation is:

r = R E cos

(EQ 8)

4.2.2 Now calculate the tangential velocity and centripetal force on a person of mass M=75 kg on the actual Earth for all three above locations using EQ 5 and EQ 6.

4.2.3 Theorize on the physical effects of the centripetal force on humans. (HINT: is the centripetal force at any of these three latitudes significant, compared to the force of gravity?).

4.3 Real Life Example: The Centrifuge


Clearly forces do work; so, centripetal forces should be no exception. An example of the use of centripetal forces doing work under extreme conditions is the centrifuge. A centrifuge is simply a device which spins at extremely high rates of speed. Since force depends upon mass, liquids (and in very extreme cases, solids) in a centrifuge can be separated into their component parts if those parts have different masses. Given a centrifuge with a radius of 20 cm and a speed of 60,000 RPM (Rotations Per Minute):

11

EXP IV

4.3.1 Calculate the force exerted on an amoeba of mass m = 1x10-8 kg in the centrifuge.

4.3.2 How many times the force of gravity on the amoeba is this? Show all your calculations.

EXP IV

12

Name: Partners name: Date and time:

Physics 135AL

Experiment 5 Collisions in one dimension OBJECTIVE To study the momentum and energy in one-dimensional collisions. EQUIPMENT Dynamics track (2.2 m, PASCO), bulls-eye level (CRAFTSMAN), collision carts (2), cart mass (0.5 kg), motion sensors (2), 750 USB (or 700 SCSI) interface, and computer with DataStudio software. INTRODUCTION The nearly elastic collisions between two carts of equal or different masses will be studied and analyzed in real time. The motion sensors will detect the instantaneous positions of the two carts. Using this information the momentum and kinetic energy of the carts will be plotted as a function of time. The corresponding conservation laws will be inspected. BACKGROUND Collisions between two (or more) objects provide a good case to study both the conservation of momentum and conservation of kinetic energy. The former is the conservation of a vector and the latter a scalar quantity. The change of the momentum of an object is given by the time integral of the force acting on the object. From Newtons third law, i.e. actionreaction principle, the forces exerted by two objects on each other during a collision are equal and opposite. Therefore the changes of momentum of the two objects are equal and opposite. Then if there are no external forces acting on the objects, the total momentum of the two objects remain the same. The conservation of the kinetic energy requires that the force exerted by the two objects on each other during the collision must be conservative. In mathematical terms this means that the force is related to a potential energy. Before the collision, lets take the potential energy to be zero. Then, if we assume that the force of the collision is repulsive, the potential energy will increase when the two objects get close to each other (the start of the impact), will reach a maximum when they are closest, and decrease and go back to zero when they are away from each other. Therefore the total potential energy doesnt change during the collision. The change in the kinetic energy is the opposite of the change in the potential energy. Therefore the kinetic energy will do the exact opposite of the potential energy: it will decrease during the impact, and it will return to its original value when the two objects are away from each other. Note that we assumed that all the internal forces that are involved in the collision are elastic and there are no external forces. Collisions satisfying these two conditions are known as elastic collisions, in which the total kinetic energy is conserved (except during the impact).

Experiment 5 5-2

The notation that we will be using is as follows:

x1 and x 2 represent the individual positions of the two objects. p1 = m1 v1 , p2 = m1v 2 , and P = p1 + p 2 represent the individual momentums and total momentum.
1 1 2 m1v12 , K 2 = m2 v2 , and K = K1 + K 2 represent the individual kinetic energies and total 2 2 kinetic energy. K1 = Note which quantity is a vector and which a scalar. Since this experiment is in 1-D, we will represent vector quantities as signed real numbers. Therefore from this point on, our notation will show v and p as v and p . However one should be cautioned that v and p are not just the magnitude (absolute value) but they also indicate direction, with the left direction corresponding to a negative value and the right a positive value. If we denote the initial and final states by the indices i and f , the conservation equations take the following form in 1-D:
i m1v1i + m2 v2 = m1v1f + m2 v2f and 1 1 1 1 i m1v1i 2 + m2 v2 2 = m1v1f 2 + m2 v2f 2 . 2 2 2 2

(1) (2)

Given the masses and initial velocities, these two equations can be solved for the two final velocities. But it involves painful algebra! Fortunately there is a much easier way, if we use the beauty of physical insight. This will be described next.
Center-of-mass frame

The center-of-mass reference frame is a reference frame that moves with the center-of mass velocity vCM with respect to the laboratory reference frame. In the center-of-mass reference frame, the velocities are given by

v1CM = v1 vCM etc.

(3)

Here v1 is the velocity in the laboratory frame and v1CM is the velocity in the center-of-mass frame. The center-of-mass frame is defined such that the total momentum in the center-of-mass frame is zero. Then we have
CM CM P CM = p1CM + p2 = m1v1CM + m2 v2 = m1 ( v1 vCM ) + m2 ( v2 vCM ) 0 .

(4)

Solving this gives

Experiment 5 5-3

vCM =

m1v1 + m2 v2 P , = m1 + m2 M

(5)

with M and P being the total mass and total momentum, respectively. From the definition of the center-of-mass frame,
CM P CM = p1CM + p2 0 ,

(6)

comes the following consequence:


CM p2 = p1CM .

(7)

This makes the solution of the conservation laws trivial. If you substitute it into the equation for the conservation of kinetic energy, it results in
iCM fCM p1,2 = p1,2 (for either object).

(8)

And thats the entire final answer in the center-of-mass frame. The plus sign simply corresponds to no collision at all. If there is a collision in the center-of-mass frame, all we have is that the velocities of both objects reverse in direction, with the magnitude unchanged. How simpler than this could it get?
Going back to the laboratory frame

We could now immediately use Equations 3 and 5 to go back to the laboratory frame. Substituting these equations into v1fCM = v1iCM (Equation 8) easily results in
v
f 1

( m1 m2 ) v1i + 2m2v2i . =
m1 + m2

(9)

i Similarly if we substitute Equations 3 and 5 into v2fCM = v2CM , we instantly obtain

f 2

( m2 m1 ) v2i + 2m1v1i . =
m1 + m2

(10)

Hence the center-of-mass frame enabled us to derive these complicated equations in the laboratory frame with no hardship.
PROCEDURE

Although this experiment is fairly straightforward, it requires you to be meticulous in order for it

Experiment 5 5-4

to work well. First you need to prepare the apparatus as described below.
Leveling the track

The track needs to be perfectly level, which you should achieve as follows. There are four feet attached to the track, with screw threads. Check to see if one pair of feet is set near the 75-cm mark and the other at the 150-cm. Correct otherwise. Then turn the knob of each four feet so that the screw thread is approximately at the vertically middle position. Next put the bulls-eye level directly above the left feet and adjust the twist of the track, meaning that the bubble should be centered up and down. Repeat it for the right feet. Then put the bulls-eye level near the 115-cm mark and carefully adjust the feet on either side so that the bubble is centered left and right. The track might be slightly bent; therefore, check the leveling for various positions and find the best compromise. Note that if you move the track on the laboratory bench, you need to level it again, since the laboratory bench is not uniform. Also make sure that the track is free of any dust.
Preparing the carts

The axes should be centered so that the wheels have the same clearance from the left and the right. Check this by holding the cart in your hands and looking at the bottom of the cart. Correct the alignment by pushing on the wheels with your thumbs from the outer side. Place the left and right carts on the track, with their non-Velcro ends facing each other. Check if each cart moves smoothly. Replace it if there is a problem.
Preparing the motion sensors

There should be a motion sensor attached to either end of the track. The motion sensors simply slide in and clamp onto the track. The beam switch on the sensor should be at the narrow setting. Align the angle of each sensor to 0 using the large knob on its side. The motion sensors are the most critical part of this experiment. If they are not properly aligned or set, you will not obtain meaningful results. The realignment of the angles of the motion sensors might be necessary if they dont record the data properly.
DataStudio

With the 750 USB (or 700 SCSI) interface turned on, start DataStudio. If you are using the 700 SCSI interface, it should be turned on before the computer is turned on; otherwise, you will need to restart the computer. In DataStudio choose Create Experiment. If this window doesnt appear, click Setup. The

Experiment 5 5-5

Experiment Setup window will show you the interface.


Click on Channel 1 on the interface and add a Motion Sensor. Click on Channel 3 on the interface and add a second Motion Sensor. In the left column under Data, you should see Position, Ch 1&2 and Position, Ch 3&4. If you see the wrong channels, delete the sensors and repeat the process. In the Experiment Setup window, change the Sample Rate to 20 Hz. This simply means that the computer will collect 20 data points, equally spaced in time, every second. Now make sure that the yellow and black plugs of the left motion sensor are connected to Channels 1 and 2, respectively. Then make sure that the yellow and black plugs of the right motion sensor are connected to Channels 3 and 4, respectively.

The motion sensor sends out an ultrasonic (of frequency greater than what we can hear) sound burst for each data point. This burst of sound will reach the cart and be reflected back to the motion sensor. The round-trip time between the sensor and the cart for this sound burst is measured by the sensor. Then the computer easily calculates the distance of the cart from the motion sensor by using the known speed of sound in air at room temperature. The speed of sound in air is 345 m / s at 73 F and 343 m / s at 68 F , changing proportional to the square root of the absolute temperature in Kelvin ( = Celsius + 273.15 ). The computer also calculates the velocity by dividing the change in the position for two adjacent data points by the change in the time, i.e. v = x / t , with t = 1/ Sample rate , which is 0.05 s for 20 Hz . You can plot Data in DataStudio in various ways by dragging a Data measurement in the left column onto some appropriate Display, such as Graph. But the real power of DataStudio is its ability to Calculate formulas. In this experiment we are interested in momenta and kinetic energy; therefore, lets set up these variables.
Setting up the constants Mass

Click the Calculate icon in DataStudio. Click the + sign next to Experiment Constants in the Calculator window. Click New under Experiment Constants. Change the name of the constant to m1. Weigh your left cart on the electronic scale and enter the mass in kg in Value. Enter kg in Units.

Experiment 5 5-6

Click Accept under Experiment Constants. Click New under Experiment Constants. Change the name of the constant to m2. Weigh your right cart on the electronic scale and enter the mass in kg in Value. Enter kg in Units. Click Accept under Experiment Constants.

Setting up the variables Position

In the Calculator window, enter x1=position12 in Definition. Click Accept at the top. Click on the triangle icon (pull-down menu) under Variables. Choose Data Measurement. Choose Position, Ch 1&2. Click Properties. Make sure Y is selected under Variable Name. Change Y to x1. Type m in Units. Under Variable Name, using the triangle icon (pull-down menu), select Time. Change Time to t. Make sure Units is s. Make sure Time is selected under Type. Click OK. Click Accept at the top of the Calculator window. Click New at the top of the Calculator window. Enter x2=-position34 in Definition. (Note the minus sign.) Click Accept at the top. Click on the triangle icon (pull-down menu) under Variables. Choose Data Measurement. Choose Position, Ch 3&4. Click Properties. Make sure Y is selected under Variable Name. Change Y to x2. Type m in Units. Under Variable Name, using the triangle icon (pull-down menu), select Time. Change Time to t. Make sure Units is s. Make sure Time is selected under Type. Click OK. Click Accept at the top of the Calculator window.

Experiment 5 5-7

Velocity

Click New at the top of the Calculator window. Enter v1=velocity12 in Definition. Click Accept at the top. Click on the triangle icon (pull-down menu) under Variables. Choose Data Measurement. Choose Velocity, Ch 1&2. Click Properties. Make sure Y is selected under Variable Name. Change Y to v1. Type m/s in Units. Under Variable Name, using the triangle icon (pull-down menu), select Time. Change Time to t. Make sure Units is s. Make sure Time is selected under Type. Click OK. Click Accept at the top of the Calculator window. Click New at the top of the Calculator window. Enter v2=-velocity34 in Definition. (Note the minus sign.) Click Accept at the top. Click on the triangle icon (pull-down menu) under Variables. Choose Data Measurement. Choose Velocity, Ch 3&4. Click Properties. Make sure Y is selected under Variable Name. Change Y to v2. Type m/s in Units. Under Variable Name, using the triangle icon (pull-down menu), select Time. Change Time to t. Make sure Units is s. Make sure Time is selected under Type. Click OK. Click Accept at the top of the Calculator window.

Momentum

Click New at the top of the Calculator window. Enter p1=m1*v1 in Definition. Click Accept at the top. Click Properties. Under Variable Name, using the triangle icon (pull-down menu), select Y.

Experiment 5 5-8

Change Y to p1. Type kg m/s in Units. Under Variable Name, using the triangle icon (pull-down menu), select X. Change X to t. Type s in Units. Select Time under Type. Click OK. Click Accept at the top of the Calculator window. Click New at the top of the Calculator window. Enter p2=m2*v2 in Definition. Click Accept at the top. Click Properties. Under Variable Name, using the triangle icon (pull-down menu), select Y. Change Y to p2. Type kg m/s in Units. Under Variable Name, using the triangle icon (pull-down menu), select X. Change X to t. Type s in Units. Select Time under Type. Click OK. Click Accept at the top of the Calculator window.

Kinetic energy

Click New at the top of the Calculator window. Enter K1=m1*v1^2/2 in Definition. Click Accept at the top. Click Properties. Under Variable Name, using the triangle icon (pull-down menu), select Y. Change Y to K1. Type J in Units. Under Variable Name, using the triangle icon (pull-down menu), select X. Change X to t. Type s in Units. Select Time under Type. Click OK. Click Accept at the top of the Calculator window. Click New at the top of the Calculator window. Enter K2=m2*v2^2/2 in Definition. Click Accept at the top. Click Properties. Under Variable Name, using the triangle icon (pull-down menu), select Y.

Experiment 5 5-9

Change Y to K2. Type J in Units. Under Variable Name, using the triangle icon (pull-down menu), select X. Change X to t. Type s in Units. Select Time under Type. Click OK. Click Accept at the top of the Calculator window.

Total momentum and total kinetic energy

Click New at the top of the Calculator window. Enter P=p1+p2 in Definition. Click Accept at the top. Click Properties. Under Variable Name, using the triangle icon (pull-down menu), select Y. Change Y to P. Type kg m/s in Units. Under Variable Name, using the triangle icon (pull-down menu), select X. Change X to t. Type s in Units. Select Time under Type. Click OK. Click Accept at the top of the Calculator window. Click New at the top of the Calculator window. Enter K=K1+K2 in Definition. Click Accept at the top. Click Properties. Under Variable Name, using the triangle icon (pull-down menu), select Y. Change Y to K. Type J in Units. Under Variable Name, using the triangle icon (pull-down menu), select X. Change X to t. Type s in Units. Select Time under Type. Click OK. Click Accept at the top of the Calculator window.

Setting up the graphs

We will produce a Graph display, which will display all the variables that we will be measuring. Do this as follows.

Experiment 5 5-10

Generate an x1 vs. t graph by dragging x1 under Data in the left column onto Graph under Displays in the left column. Enlarge (maximize) the Graph 1 window, which has just been generated. Then click on x2 under Data in the left column but dont release the mouse. Drag x2 into the middle of the Graph 1 window and release the mouse only after the entire graph, not one of the axes, is highlighted. If you release the mouse with one of the axes highlighted, x2 will replace the variable on that axis, in which case you need to delete Graph 1 and start the whole process over. Repeat the above process of dragging the variables for v1, v2, p1, p2, K1, K2, P, and K. If you release the mouse incorrectly in the process and replace one of the previous variables with the new variable, delete Graph 1 and start over. When you are done, the graph should look like this:

Figure 1. DataStudio setup.

Experiment 5 5-11

If there are missing variables or if there is another variable on the time axis, you released the mouse at the wrong place. If X shows instead of t on the time axis, you forgot to change the time argument for either p1, p2, K1, K2, P, or K under Properties in the Calculator window. If time shows instead of t on the time axis, you forgot to change the time argument for either x1, x2, v1, or v2. If Y shows on the y-axis, you forgot to change the y-argument for one of the variables. In any case you need to resolve the problem before you can take data.
Taking data CAUTION: Make sure that the carts never reach and hit the motion sensors. Stop the carts with your hands before they reach the motion sensors. Case 1: m1 = m2

Place the left and right carts on the track, with their non-Velcro ends facing each other. Make sure that in the Calculator window, you entered m1 and m2 for the left and right carts correctly and clicked Accept after entering each. If not, correct the problem and click Accept at the top of the Calculator window to finish the calculation at the end. Record the values for m1 and m2 on the laboratory-report-summary page. First make a trial run. Click the Start button and move both carts by hand. Make sure that your hands dont interfere with the ultrasonic beams of the motion sensors. See if the motion sensors record the data correctly and smoothly. If there is a problem, check the alignment of the motion sensors and check whether there is something in the view of the motion sensors, such as cables, foreign objects on the laboratory bench, your hands, etc. When you are satisfied, click the Stop button. Then make sure to select Delete ALL Data Runs under the Experiment menu so that you will not display useless data on your graphs. Now it is time to make a clean data run. You will be in charge of pushing the left cart and also catching the carts before they hit the motion sensors. Your partner will be in charge of controlling the Stop button and he/she will help you catch the carts before they hit the motion sensors. Put the right cart near the middle of the track and the left cart at about 20 cm from the left motion sensor. Both carts should be stationary at this point. Now click the Start button. Give the left cart a controlled push so that the carts neither go too fast to bounce off the track or touch each other during the collision, or too slow for the friction to become very significant. Your hand shouldnt get in the way of the motion sensors when you push or stop the carts. Your partner should click the Stop button before one of the carts reach the end of the track. You and your partner should make sure to catch the carts before they can hit the motion sensors. Now you should adjust the scales of your graphs. Click on each four graph and do an Autoscale using the icon. Then position the mouse over the time axis and adjust the

Experiment 5 5-12

time axis so that the timeframe near the collision event is displayed. When the mouse pointer changes to a hand symbol, you can move the axis. When it changes to a spring symbol, you can shrink or expand the axis. When you are done, your graphs should like the example below. Otherwise Delete ALL Data Runs under the Experiment menu, make

Figure 2. Typical screen shot for the data and analysis in the equal-mass case.

Experiment 5 5-13

a new run, and scale your graphs again. You will record all your quantitative answers on the laboratory-report-summary page, and for each quantity, you should also include the error. You will determine the error through a linearregression fit by the software. How to do this will be described in Question 4.
Question 1. Is the total momentum conserved throughout the data run?

Question 2. Is the total kinetic energy conserved throughout the data run?

Question 3. What is happening to the kinetic energy at the collision instant?

Question 4. On the graph for the kinetic energy, click on Run #1 for the total kinetic energy K in the legend of the graph. Make sure to select the correct curve (K, not K1 or K2); otherwise, this will not work. Select the data points when the left cart is moving with constant velocity (the initial state) using your mouse. The selected data points should show highlighted in yellow. Then

do a linear fit using the

icon. Your fit should look very similar to the picture above.

Now using the xy tool , determine the midpoint of your collision and the extrapolated total kinetic energy before the collision (the initial state). The extrapolated total kinetic energy is the y-value of the linear fit at the midpoint of the collision, as read by the xy tool. Record the initial K in the laboratory-report-summary page. Read the error from the linear-regression fit (rootmean-square [rms] error) and report this in the summary page as well. In the example above, the values for the initial K are 0.106 0.001 J . Repeat the procedure for after the collision (the final state). In the example above, the values for the final K are 0.104 0.003 J . Note: In order to delete an unwanted fit, click on the variable, for which you want to delete the fit, in the legend. Then unselect the fit or select No Curve Fits, using the Fit icon.
Question 5. Determine the initial and final values for v1 , v2 , p1 , p2 , P , K1 , and K 2 , as well as the errors in these quantities using the same analysis as in Question 4:

First click on the variable in the legend of the graph in DataStudio.

Experiment 5 5-14

Then do a linear fit for the initial state. Use the xy tool to determine the quantity from the extrapolation of the linear fit at the midpoint of the collision. Record the error from the linear-regression fit. Repeat this for the final state.

In the laboratory-report-summary page, fill out the table with the information you obtain from your graphs. Also fill out the P and K information and comment on whether the momentum and kinetic energy are conserved within the errors. Note that when you calculate P = P2 P , 1
2 the absolute errors would combine as = 12 + 2 .

Question 6. Using the measured value for the initial v1 , calculate the final v1 and final v2 using the theoretical formulas. Report these in the laboratory-report-summary page along with the percentage errors for the measured values with respect to these theoretical values.

Obtain a printout so that you can include it in your write-up. Save the DataStudio file under the name equalmass.

Case 2: m1 > m2

Now save the DataStudio file under the name largermass. Under the Experiment menu, choose Delete ALL Data Runs. Put the 0.5 kg cart mass on the left cart and measure its mass with the cart mass, using the electronic scale. In DataStudio open the Calculator window (Calculate) and change m1 to the new value in kg, including the cart mass. Click Accept next to the Experiment Constants. Then click Accept at the top of the Calculator window to finish the calculation.

Experiment 5 5-15

Record the values for m1 and m2 on the laboratory-report-summary page.

Question 7. Is the total momentum conserved throughout the data run?

Question 8. Is the total kinetic energy conserved throughout the data run?

Question 9. Do the analysis as in Question 4. Question 10. Do the analysis as in Question 5. Question 11. Using the measured value for the initial v1 , calculate the final v1 and final v2 using the theoretical formulas. Report these in the laboratory-report-summary page along with the percentage errors for the measured values with respect to these theoretical values.

Obtain a printout so that you can take include it in your write-up. Save the DataStudio file under the name largermass.

Case 3: m1 < m2

Now save the DataStudio file under the name smallermass. Under the Experiment menu, choose Delete ALL Data Runs. Switch the left cart (with the mass) with the right cart. In DataStudio open the Calculator window (Calculate) and switch the values of m1 and m2. Click Accept both for m1 and m2. Then also click Accept at the top of the Calculator Window to finish the calculation. Record the values for m1 and m2 on the laboratory-report-summary page.

Experiment 5 5-16

Question 12. Is the total momentum conserved throughout the data run?

Question 13. Is the total kinetic energy conserved throughout the data run?

Question 14. Do the analysis as in Question 4. Question 15. Do the analysis as in Question 5. Question 16. Using the measured value for the initial v1 , calculate the final v1 and final v2 using the theoretical formulas. Report these in the laboratory-report-summary page along with the percentage errors for the measured values with respect to these theoretical values.

Obtain a printout so that you can take include it in your write-up. Save the DataStudio file under the name smallermass.

Experiment 5 5-17

Physics 135ALphysics for the life sciences Laboratory-report summary

Experiment 5 Collisions in one dimension


Name: ____________________________ Partner: __________________________ Your lecture instructor: _____________ (not your partners)
Table 1. Masses.

Date: ___________________________ Lab period: ______________________ Lecture period: ___________________

m1 = m2
Left cart Right cart

m1 > m2 m1 = _._ _ _ _ kg m2 = _._ _ _ _ kg

m1 < m2 m1 = _._ _ _ _ kg m2 = _._ _ _ _ kg

m1 = _._ _ _ _ kg m2 = _._ _ _ _ kg

Table 2. Experimental data and error.

m1 = m2
Initial Final Initial

m1 > m2
Final Initial

m1 < m2
Final

v1 v2
v1theory final
theory v2 final

% error = % error =

% error = % error =

% error = % error =

p1 p2
P P

K1 K2
K K
K K 100

You must complete: This summary page and your answers to Questions 116.

Experiment 5 5-18

Name: Partners name: Date and time:

PHYS135AL

EXPERIMENT VI ROTATIONAL KINEMATICS

1.0 Objective
To gain both a mathematical and physical understanding for the moment of inertia, to examine conservation of energy in systems that include rotational motion.

2.0 Background
In the experiments you have done so far you have studied bodies that have performed some motion in space, also called translational motion. But an extended body, like a cylinder or a ball, can also rotate around one of its body axes. If it is doing so, there has to be some energy associated with this rotation, since all the atoms of the body are in more or less rapid motion. How can we determine the amount of energy that is necessary to make a given body rotate about a given axes with a given rotational speed? And how does the amount of energy depend on the shape of the body?

2.1 The moment of inertia


It is this last question which is best answered by a physical property of every body with mass, known as the moment of inertia, I. For an object that is composed of several shapes with a common axis, the total moment of inertia is the sum of the moments of inertia of its parts. So the moment of inertia, I, of a body with mass m = is defined to be the sum:

mi

about an axis r

I =

mi ri
i

(EQ 1)

where ri is the distance from the axis to the ith small mass mi. 2.1.1 Can I ever be less than zero or negative? Explain why or why not.

2.1.2 What is the moment of inertia for a point particle of mass m a distance R from an axis?

EXP VI

In english, the moment of inertia of a body with mass measures its tendency to want to rotate around any axis you wish when a force is applied to it. It is a physical property of each individual object, and depends upon density, shape, size, and the axis about which the moment of inertia is being calculated. 2.1.3 What are the units of the moment of inertia?

2.1.4 Which has the greatest moment of inertia: a 2 kg point-particle at a distance of 1 m from an axis, or a 1 kg point-particle a distance of 2 m from the same axis? Explain.

The real world consists of objects which are not point-particles. It takes calculus to determine moment of inertia for extended objects, but if the body is a simple solid shape then the moments of inertia are fairly easy to calculate. Note that the total moment of inertia for a system of extended objects is just the linear sum of the individual parts moments.

Figure 1: A ring and a cylinder of radius R and equal mass.


Thin Ring Solid Cylinder

ri mi

ri

mi

For a thin ring of radius R and total mass M, the answer is Ithin ring = M R2 since all ri are equal to R. 2.1.5 For a solid cylinder of the same mass as the ring, will the moment of inertia be smaller than this or larger? Explain. (hint -- see EQ 1).

EXP VI

The result for a solid cylinder of mass M and radius R is

MR I cylinder = ---------2
ring, while R2 is its inner radius.

(EQ 2)

For a thick ring (figure 2), the sum comes out to give Ithick ring = M (R12 + R22) / 2, where R1 is the outer radius of the

Figure 2: A cylinder and a thick ring of same mass M:


Solid Cylinder Thick Ring

R2 R R1

For a sphere of radius R and mass M (not shown), the moment of intertia is I sphere

2 = -- MR 2 5

2.2 Torque
As weve seen previously, if no net force acts upon a body then it is in translational (i.e. linear) equilibrium. To cause a body to rotate, a net torque,

, must be applied to it. Torques are forces applied at a distance which cause rotation. = I
(EQ 3)

Since torque is a vector, it has a direction and a magnitude associated with it. Specifically,

where I is the moment of inertia of the object and no net torque is applied, i.e.

is the angular acceleration of the object (see Experiment V). If

i =
i

0 , then a body will be in (angular) equilibrium. Thus for an object to be in true

equilibrium, the sum of the forces acting upon it and the sum of the torques acting upon it must both be zero.

2.3 Energy of rotating objects


Previously, weve looked at the energy associated with objects that were undergoing translational (i.e. linear) motion. One must determine the energy associated with rolling or rotating motion as well as its translational energy when considering the total energy of an object. The moment of intertia is integral in determining this total energy if there is rotation. The angular frequency,

= 2f (see Experiment V), of the rolling motion is also important.

The total kinetic energy of a object undergoing both translational and rotational motion is:

EXP VI

1 2 1 2 KE total = -- mv + -- I 2 2
or system:

(EQ 4)

As always, the total energy of an object or system, Etotal, is just the kinetic energy plus the potential energy of the object

E total = ( KE + PE )

(EQ 5)

Consider the following thought experiment -- imagine you have a sphere, a cylinder and a thin ring all of equal mass and equal outer radii. 2.3.1 If you were to race these objects down a slope from rest, which object do you think would win?

2.4 Momentum of rotating objects


Clearly since weve defined energy associated with rotation, we can also define a rotational momentum. The angular momentum, L, of a rotating object is just its angular frequency multiplied by its moment of inertia.

L = I
In a system where the sum of the torques is zero,

(EQ 6)

i =
i

0 , the systems angular momentum is conserved:


(EQ 7)

L initial = L final

3.0 Procedure
In the first part of this experiment, you will see the effect of a changing radius on the moment of inertia of a person seated upon an essentially-frictionless stool. TAKE ALL MEASUREMENTS IN MKS UNITS.

3.1 Conservation of angular momentum


Have one partner sit on the frictionless stool holding a 1 kg weight in each hand. 3.1.1 Record the length of one arm, from the point it connects with the shoulder to the middle of the 1 kg weight.

EXP VI

With the weights held at arms length and at shoulder height, begin rotating the stool. Pick one of the weights as your zero point for counting revolutions, and use a timer to record how long it takes to complete a revolution. 3.1.2 Determine the initial period, T1, for a full revolution, counting from your zero point mass.

Now have the person slowly draw their arms in as close to their bodies as possible, making certain the weights are pulled in keeping their height above the ground constant. 3.1.3 Determine the final period, T2, for a full revolution, counting from your zero point mass.

After taking data, you can now analyze it and determine the moment of inertia: 3.1.4 Determine 1, the angular velocity of the person-mass system with arms extended, from your value of T1.

3.1.5 Determine 2, the angular velocity of the person-mass system with arms drawn in, from your value of T2.

For the person-mass system, we can consider the stool to be essentially frictionless. Thus no external forces or torques acted upon the system and angular momentum should be conserved:

L 1 = L 2 , which means I 1 1 = I 2 2

I 1 can be considered the sum of the bodys moment of inertia and the moment of inertia of the two masses (well
ignore the weight of the arms and hands). I 2 is the inertia of the body and weights together, which is essentially the same as the bodys moment of inertia alone. Making these approximations and doing some math, we find that

EXP VI

1 I body = ------------------ 2mr 2 2 1

(EQ 8)

where m is the mass of one of the hand-held weight and r is the measured distance from the center of chest to the mass (with arms stretched out). 3.1.6 Using EQ 8, determine the moment of inertia of the person on the stool. Dont forget units.

3.1.7 Approximate the person on the stool as a cylinder, and use EQ 2 to independently determine the persons moment of inertia.

3.1.8 Compare and contrast the two values of I and theorize which method is best.

3.1.9 In this experiment the energy is NOT conserved! Why? (Hint: Is there any work done in this experiment?)

EXP VI

3.2 Back to the races


You should now have a photogate timer, an accessory photogate, a solid cylinder, a hoop, and a wooden ramp. First have someone place their coat or a towel at the base of the ramp so that objects which roll down are stopped safely. 3.2.1 Weigh both objects and measure their radii, R, recording each in your data sheet.

3.2.2 Test your response to 2.3.1. Which object wins? Race the objects three times to test the consistency of the result.

3.2.3 Does the fact that one object has a greater moment of inertia completely explain your results in 3.2.2? Why or why not?

We will now experimentally determine the moments of inertia of both objects. Set the photogate timer to PULSE mode and set up the photogate timer and accessory photogate with the ramp as illustrated in figure 3:

Accessory photogate

Photogate timer

Ramp

coat or rags

Figure 3: ramp and photogate set-up

EXP VI

Place the accessory photogate so it hangs over the ramp as close to the top as possible. Place photogate timer so that it hangs over the ramp as near the bottom as possible (but not past the bottom). The photogate timer in PULSE mode will measure the time interval between when the light beams of the two photogates are broken. Adjust the height of the photogates so this interval actually corresponds to the time it takes for the hoop to travel down the track. (Why is this important?) When everything is adjusted, you should be able to release the hoop from rest and have the photogates measure the time it takes for the hoop to roll down the entire ramp. During the experiment you have to release the cylinder and the hoop as close as possible to the trigger point of the accessory photogate. Otherwise they already have a significant velocity before entering the accessory photogate. When the setup is ready to go, measure the distance along the ramp L and height H traveled by the object in the following way: First mark the two contact points the object would make on the ramp just before it is released and just before it is detected by the sensor of the photogate timer at the bottom. Then measure the distance along the ramp and difference of height between these two points. 3.2.4 Record L and H in your data sheet.

3.2.5 Measure the time it takes the hoop to roll from rest down the ramp three times.

3.2.6 Calculate the average time.

3.2.7 Repeat the same process for the cylinder.

EXP VI

3.2.8 For both the hoop and the cylinder, calculate their average velocities vav, using L and the average times you calculated above.

Since the acceleration along the track is constant, the final velocity vf, when the body leaves the ramp, will be twice the average velocity, vf=2vav. Let us now derive a formula to determine the moment of inertia of the cylinder and the hoop using the measured value of the average velocities. Consider therefore the energy balance for this experiment. Before each body is released, it has zero kinetic energy and a potential energy given by:

PE = mgH

(EQ 9)

This potential energy is the initial energy of the body. Upon reaching the second photogate, the body has lost the potential energy, but now has a final kinetic given by EQ 4. Conservation of energy for this experimental set-up says that Einitial = E final,or PE = KE. 3.2.9 Use EQ 4 and EQ 9 to solve for the moment of inertia of the body. You will also have to use EQ 5 from experiment V. In the end, express your equation in terms of vav. HINT: You should end up with:

2 gH I exp = mR --------- 1 2v 2 av

(EQ 10)

3.2.10 Use EQ 10 to obtain the experimental values of the moment of inertia for both the cylinder and the hoop.

EXP VI

3.2.11 Calculate the theoretical values for each objects moment of inertia (see section 2.1) and compare them to your answer in 3.2.10

3.2.12 What are sources of error for this experiment? Determine their effect on the value of Iexp.

EXP VI

10

Name: Partners name: Date and time:

PHYS 135AL

Experiment 7 Forces and torques in static equilibrium I: cantilever OBJECTIVE To understand the distribution of forces and torques in an extended body maintained in static equilibrium. To learn how to use graphs to analyze experimental data. EQUIPMENT Meterstick; sliding hangers (3); U hangers (2); motion sensor; force sensors (2); clay-bumper assembly, including the clay-cup attachment, 1-in-long clay, and penny with the clay dot; hook attachment; distance-measure sleeve; weight attachment (500 g); two vertical rods clamped on the table; horizontal rod clamped between the vertical rods; small rod to attach the motion sensor to one of the vertical rods; ruler (50-cm); and computer, interface, and DataStudio software. INTRODUCTION In this experiment you will use the DataStudio software to measure and analyze the forces and torques on a meterstick cantilever system. Force sensors will measure the forces at the attachment points. A motion sensor will be used to measure all the distances involved. You will use the DataStudio software to plot one of the forces against the distance of the hanging weight, of which you will then analyze to calculate the hanging weight, weight of the meterstick, and the center-of-mass of the meterstick. THEORY An extended body said to be in static equilibrium should have no linear or rotational motion. (Strictly speaking no linear or rotational acceleration is sufficient.) These conditions can be satisfied by two equations:

F = 0
i i

Equation 1 Equation 2

=0

In fact these equations are very simple and just state that the sum of all forces and sum of all torques should equal to zero. Forces are vector quantities, and you need to separate them into their horizontal and vertical components before you can do arithmetic with them. Our cantilever system is shown in Figure 1. A meterstick hangs horizontally from two force sensors positioned vertically on a horizontal support bar. A weight is attached to the other side of the meterstick. In addition the meterstick with sliding hangers has its own weight, which can be thought as a single weight acting on the center of mass of the metersticksliding hangers system.

Experiment 7 7-2

support bar Force sensor a Force sensor b

clay cup clay penny sliding hangers

Fa
hook

Fb
U hanger CM meterstick sliding hanger

Ws
motion sensor

U hanger distance-measure sleeve weight attachment

db

ds

Figure 1. Meterstick cantilever system. The meterstick hangs from two force sensors via two sliding hangers. A weight is attached on the other side with another sliding hanger.

The forces in this experiment act only along the vertical direction. Therefore the vectors will be just signed numbers. Then Equation 1 becomes

Fa + Fb + Ws + W = 0 .

Equation 3

In order to write down the torque equation, we need to define an origin. You can in fact choose any point that is convenient to you. Lets choose Force sensor a as the origin. Remembering that torque is perpendicular distance between the line of force and origin times force, we have

0 Fa + db Fb + d s Ws + d W = 0 .

Equation 4

For the purposes of this experiment, we rearrange Equations 4 and 5 so that we end up with

Ws = Fa Fb W and
W d W Fb = d + s s . db db

Equation 5 Equation 6

Experiment 7 7-3

The quantities in these equations are as follows.

db : the distance of Force sensor b, d s : the distance of the center of mass, where the center of mass is that of the meterstick and two sliding hangers the force sensors attach to,
d : the distance of the weight attachment,

Fa and Fb : the forces applied on the meterstick by the force sensors, Ws : the weight of the meterstick, including the two stationary sliding hangers (one of them with a U hanger) but excluding the moving sliding hanger,
W : the weight of the weight attachment, including the sliding hanger, U hanger, and distance-measure sleeve.

In this experiment we will move the weight attachment along the meterstick; therefore, d is a variable quantity. Since the sliding hanger, U hanger, and distance-measure sleeve will also move along the meterstick, we include their weight in W as well. Then the two stationary sliding hangers must be included in the weight of the meterstick. Therefore d s represents the center of mass of the meterstick and two stationary sliding hangers as a whole. The equation for a straight line plotted in the xy-plane is given by

y = mx + b ,

Equation 7

where m is the slope and b is the y-intercept. Then Equation 6 represents a line with the slope and y-intercept given by
W and db d W b= s s . db m=

Equation 8 Equation 9

If we make a plot of Fb versus d , we can calculate W and d s from the slope and y-intercept.

Experiment 7 7-4

Equipment setup Refer to Figure 1 and see if everything is set up properly. CAUTION: To install the distance-measure sleeve, push it gently straight up from below. And pull it gently straight down from below to remove it. The small holes on the clamps inside the distance-measure sleeve snap in and out automatically on the protrusions of the sliding hanger. Do not put your fingers inside the distance-measure sleeve to position the clamps, which may be damaged by doing so. 1. Adjust the horizontal support rod so that the two force sensors are exactly at the same height. (Use the 50-cm ruler.) 2. Force sensor a should be about 20 cm from the motion sensor. 3. Force sensor b should be about 20 cm from Force sensor a. 4. The motion sensor should be about 15 cm below the meterstick. The large knob on the back of the motion sensor should be adjusted to 0 . Also check if the beam-width switch on the sensor is at narrow. And make sure that the motion sensor points straight ahead along the meterstick. 5. The hook on the force sensor should be slightly loose so that it can align with the U hanger. 6. The clay-bumper assembly has a penny at the bottom. At the bottom of the penny, there is a small dot of additional clay. The clay acts as a bumper and positions Force sensor a, also reducing vibrations. The penny prevents the screw from sinking in the clay. The small dot of additional clay under the penny prevents the screw from sliding. Tighten the screw on the sliding hanger. Then position the screw under the center of the penny. 7. Loosen the screw on the sliding hanger attached to the hook of Force sensor b. Check to see if the meterstick is horizontal, not tilted around its axis, and pointing straight ahead from the motion sensor. (Look from the front and side.) Check to see if Force sensors a and b are vertically straight. Make the necessary corrections. Then tighten the screws of Force sensors a and b. Finally tighten the screw of the sliding hanger attached to the hook of Force sensor b. Note that the sliding hangers should be centered with the force sensors. DataStudio setup DataStudio is a powerful software that allows you to collect and analyze data. Data collection is achieved by various sensors connected to a computer interface. Set up DataStudio as follows. 1. Start DataStudio and choose Create Experiment. If this option doesnt come up, click on Setup. 2. In the Experiment Setup window, under Add Sensor, double-click on the motion sensor and double-click twice on the force sensor (regular force sensor, not the student force sensor). (Note that the motion and force sensors are under Digital and Analog submenus, respectively.) Now the window should show one motion sensor connected to digital channels 1 and 2 and two force sensors connected to analog channels A and B on the interface. Check to see if the actual sensors are connected exactly in the same way.

Experiment 7 7-5

3. Click Calculate. Click Experiment Constants. Click New under Experiment Constants. Type d0 as the name of the experiment constant. Click Accept next to the name. 4. Click New under Experiments Constants again. Type db and click Accept. Then click New, type W, and click Accept. We have just defined the experiment constants that we will be using in this experiment. 5. In Definition type d=distance*100-d0. Click Accept at the top. In Variables define distance as Data MeasurementPosition. (Use the small-triangle menu icon.) d0 should have automatically been defined as the experiment constant d0. Click Properties and change Variable Name to d, Units to cm. Click OK. d will be the distance of the weight in centimeters. (100 is just the conversion factor from meters to centimeters.) d0 will be a calibration parameter that will enable us to set the distance of Force sensor a to 0, i.e., define our origin. 6. Click New at the top. In Definition type Ws=-Fa-Fb-W (i.e., Equation 5). Click Accept at the top. In Variables define Fa and Fb as Data MeasurementForce Ch. A and Force Ch. B, respectively. (Use, again, the small-triangle menu icon.) In Properties change Variable Name to Ws and Units to N (or, newton). Click OK. 7. Drag d under Data to Graph under Displays. This should generate a d versus time graph. Click the icon in the graph. This should display the mean value of d along with the minimum and maximum. 8. Drag Force Ch. B under Data to Graph under Displays. Then drag d under Data to the time axis in the graph just generated, but do not release the mouse before time axis is highlighted. This should generate a Force Ch. B versus d graph. If not, the time axis wasnt highlighted. Then just remove the graph and repeat the process. 9. As the final graph, Drag Ws under Data to Graph under Displays. This should generate a Ws versus time graph. Then click the icon in the graph to display the mean. 10. In the Experiment menu, under Set Sampling Options, check Keep data values only when commanded and uncheck all the other boxes. Click OK. This will enable manual sampling of data. Now we are set to go. Proceed with the procedure. PROCEDURE CAUTION: Make sure that the motion sensor doesnt directly see any objects other than the distance-measure sleeve, such as the force-sensor wires, items on the table, your hands, etc. Otherwise they will be detected instead. In addition make sure to properly tare the force sensors whenever needed to do so as explained in this procedure. 1. Remove the weight attachment. Then remove the U hanger on the middle sliding hanger from the hook on Force sensor b. Then slowly detach the meterstick from Force sensor a. Set the meterstick on the table. By pushing the tare button on the back, tare Force sensor a with only the clay-bumper assembly on. Tare (or, zero) sets the zero reading of the force sensor. Likewise tare (zero) Force sensor b by pushing the tare button on the back, with only the hook on. When you are done, replace the meterstick. 2. Slide the distance-measure sleeve over the sliding hanger under Force sensor a by pushing it straight up from below. See if the protrusions snapped in the holes. Wait for a

Experiment 7 7-6

few seconds until the system equilibrates and Click Start on DataStudio. Look at the d versus time graph. Click Calculate, choose the experiment constant d0, and see if it is set to zero. If not, set it to zero and click Accept. See if the data reading is stable. If it is, click Keep rapidly about 20 times. Each time you click Keep, data is recorded (hence, manual data sampling). Then click Stop. Enter the mean value of d you obtain from the graph as the value of d0 in Calculator. Click Accept. After you do that, the mean value in the graph should change to zero, within random error. Then make a new data run by clicking Start, keeping about 20 data points again, then stopping. This will generate a new curve in the d versus time graph. Look at the mean value of d in this curve and see if it is almost zero. If so, your distance calibration is done! You have defined your origin as the position of Force sensor a. If not, simply repeat this step until you obtain satisfactory results. 3. Remove the distance-measure sleeve from Force sensor a by pulling it gently straight down and replace it on Force sensor b by pushing it gently straight up. Check to see if the holes on the clamps are snapped onto the protrusions. Click Calculate and choose db under Experiment Constant. Make a new data run and collect about 20 data points. In Calculator enter the mean value from the position graph as the value of db. Click Accept. This is the distance of Force sensor b from the origin (i.e., Force sensor a). Repeat the measurements and see if you get the same result. If so, move on with the next step. 4. Now all the calibrations and initial measurements are finished and we are ready to generate the Fb versus d graph. Slide the distance-measure sleeve over the weight attachment and see if the holes on the clamps are snapped onto the protrusions. Move the weight attachment (loosen the screw if necessary) close to Force sensor b, but not touching or interfering the sliding hanger for Force sensor b. Wait until the system equilibrates. Click Start and look at the Force Ch. B versus d graph. If the data reading is fairly stable, rapidly take about 20 data points by clicking Keep. But do not click the Stop button yet. Then move the weight attachment about five centimeters further away. Wait until the system equilibrates and then rapidly Keep about 20 data points. Repeat the process until you come close to the end of the meterstick. Then click Stop. You have just finished performing the experiment. Remove the weight attachment and set it aside. ANALYSIS Go to the Fb versus d graph and make a linear fit using the Fit icon.

Question 1. What are the slope and y-intercept? slope m = y-intercept b = Question 2. Now, using the slope, solve for the weight of the weight attachment, W .

W=

Experiment 7 7-7

Enter your value for W in Calculator (click Calculate) as the value of the experiment constant W . Then go to the Ws (weight of the meterstick) versus time graph.

Question 3. From the Ws versus time graph, what is the mean value of Ws ? Is Ws almost a constant over time? If not, why?

Ws =

indicates mean value.)

Question 4. Now, using the y-intercept, solve for the distance of the center of mass of the metersticksliding hangers system, d s .

ds =
Question 5.

Now remove the meterstick but keep the two stationary hangers (one with a U hanger) on the meterstick. Generate a Force Ch. B versus time graph. Click the icon to display the mean. Start a new data run. Tare Force sensor b by pushing the tare button and see if the reading on the graph goes to zero. Then hang the weight attachment along with the sliding hanger, U hanger, and distance-measure sleeve onto the hook of Force sensor b. Record about 20 data points by clicking Keep. Stop the data run. Record the mean value for the weight of the weight attachment below. This should be the true value of the weight. Compare this value to the value you obtained using the laws of static equilibrium (your answer to Question 2). Then find the experimental error.

W actual =
Percent error =
Question 6.

W static-equilibrium laws W actual 100 = W actual

Remove the weight attachment from Force sensor b. Start a new data run. Attach a U hanger to Force sensor b and tare the sensor by pushing the tare button. Then get the weight of the meterstick and two stationary hangers (one with a U hanger) after balancing it on the hanger. Find the experimental error by comparing this value to the value you obtained earlier (in Question 3).

Wsactual =
Percent error =

Ws static-equilibrium laws Ws actual 100 = Ws actual

Experiment 7 7-8

Question 7.

Hold a U hanger with your fingers, and balance the meterstick (with the two sliding hangers still on it at the original positions) on the U hanger. The point at which the meterstick is balanced is the center of mass. Read off this value on the meterstick and read off the value for the Force sensor a sliding hanger as well (both with millimeter accuracy). Calculate the difference and record the results below. This is the actual value for the distance of the center of mass from the origin (Force sensor a). Then calculate the percent error for your experimental value (in Question 4).

d smeterstick reading =
meterstick reading da =

d sactual = d smeterstick reading d ameterstick reading = d sstatic-equilibrium laws d sactual 100 = Percent error = d sactual
Question 8. During the experiment you have observed that Fb was negative. The force sensor shows a negative force when the meterstick pulls on it, i.e., when the force sensor pulls on the meterstick (remember actionreaction forces). Therefore this means that the force Fb applied on the meterstick by Force sensor b was pointing up. Is this consistent with your expectations? What do you expect the direction of the force Fa applied on the meterstick by Force sensor a to be? Is this consistent with the fact that we were using a bumper instead of a hanger for Force sensor a? Now generate an Fa versus time graph and see if your expectations were right.

Question 9. In this experiment why did we record many data points at each distance of the weight attachment instead of just recording one data point and moving on to another distance?

Question 10. Discuss the difficulties and possible sources of error you encountered in this experiment.

Name: Partners name: Date and time:

Physics 135AL

Experiment 8 Elasticity: stress and strain OBJECTIVE To study the deformation of materials under applied force. To learn the concepts of stress and strain. EQUIPMENT PASCO stressstrain apparatus, including a rotary motion sensor and an economy force sensor; calibration bar; 3/ 8 socket with spindle bar; six different material test coupons; digital caliper; computer with DataStudio software; the beam deflection setup, including H, L, and T beams, two beam stands, and indicator gauge with stand. INTRODUCTION Test coupons made from various materials will be stretched by manually turning a screw with a crank. The amount of the stretch in the material will be measured by precisely monitoring the rotation using a rotary motion sensor. The tension force in the material, translated into a smaller force by means of a lever, will be measured using an economy force sensor. Real-time stress strain behavior for different materials will be studied in this manner. In addition the deflection of beams of various geometries under applied load will be studied using a separate setup. BACKGROUND General definitions When we studied forces and torques on an object in static equilibrium, i.e. statics, we calculated the external forces (and torques) acting on the object. From the Newtons third law, or the actionreaction principle, these external forces result in internal forces that are equal in magnitude and opposite in direction. These internal forces in the object lead to the deformations of the object, and if they are large enough, the object ultimately breaks, or, fractures. For small forces the length of the object changes by a small amount proportional to the applied force. This linear behavior could be written as

F = k L ,

(1)

where F (N) is the applied force, the proportionality constant k (N/m) is the spring constant, and L (m) is the change in length (elongation) that occurs. This is known as the Hookes law. Beyond a certain elongation, the Hookes law is no longer valid and the relationship between the force and elongation is no longer linear. The point at which the Hookes law ceases to be valid and the force vs. elongation relationship becomes nonlinear is known as the proportional limit.

Experiment 8 8-2

(See Figure 1.)


Force F (N) or stress = F/A (MPa)

Proportional limit

TS
y

Plastic region

Elastic limit Elastic region

Fracture

= 0.002
Elongation L (m) or strain = L/L (dimensionless)

Figure 1. Elongation of a material under tension. y represents the yield strength and TS the tensile strength.

When the applied force on the object is removed within the proportional limit, the object will return to its original length. This still happens to be the case until a second limit called the elastic limit. Beyond the elastic limit, the object will no longer return to its original shape when the applied force is removed. Then the object is said to enter the plastic region. Once the object is in the plastic region, permanent deformation will result. Most structures are designed so that the various components used are safe from entering the plastic region, in which the forces would be large enough to cause permanent deformation, which would in turn render the structure damaged. The spring constant k in Equation 1 depends on the original length L0 and the cross-sectional area A of the object, e.g., a steel rod. It is more practical to define a similar quantity called the Youngs modulus E, which only depends on the material the object is made of. Then for an object of any dimensions, the spring constant k can be calculated as
k = E A . L0

(2)

The larger the spring constant, the stronger (less elastic) is the spring; therefore, from Equation 2, a larger Youngs modulus, bigger cross-sectional area, and a shorter length all result in a less elastic object. By substituting Equation 2, Equation 1 can be written in terms of the Youngs modulus so that we have

Experiment 8 8-3

F L . = E A L0

(3)

Defining stress = F / A and strain = L / L0 , this could simply be put as

= E ,
i.e., stress is simply proportional to the strain if the material under stress is within the proportional limit (i.e., stress and strain are small). And the constant of proportionality is the Youngs modulus E . (See Figure 2.)

(4)

L0

Figure 2. The original length L0 and cross-sectional area A of an object under the stress result of the applied force F. The resulting elongation L corresponds to the strain

=F/A

as a

= L / L0 .

Note that the common unit for both the stress and Youngs modulus E is megapascal (MPa), which is 106 N / m 2 . The strain is dimensionless. Stressstrain characteristics of materials Figure 1 shows typical stressstrain characteristics of a metal. Of particular interest is the yield strength y , i.e. the maximum stress that the material can sustain before it deforms plastically (permanently changes it shape), or, yields. Since the definition of this point is somewhat ambiguous, it is common practice to draw a line parallel to the linear region, which crosses the xaxis at the strain = 0.002 = 2 /1000 . Then the intersection of this line with the stressstrain curve gives the yield strength y . It is important to know this value because once the yielding takes place, the material is permanently deformed and is usually considered useless thereafter. A second quantity of interest is the tensile strength TS , which is the ultimate (maximum) strength of the material. (See again Figure 1.) Some metals, especially soft steel, exhibit the so-called yield-point phenomenon. In this case the yield strength y is defined as in Figure 3. Such metals behave very linearly up to a point and then suddenly yield and enter the plastic region. Once in the plastic region, the metal starts yielding with a constant force (or stress) corresponding to the yield strength (see again Figure 3).

Experiment 8 8-4

Upper yield point Stress (MPa)

TS
y
Plastic region Lower yield point

Fracture Elastic region

Strain (dimensionless) Figure 3. Yield-point phenomenon and the definition of the yield strength y in this case. The tensile strength TS is also shown.

Another important elastic property for materials is the ductility. A ductile material is one that undergoes quite a big plastic deformation before it fractures. On the other hand a brittle material is one that shows very little or no plastic deformation upon fracture. (See Figure 4.) Quantitatively the ductility is defined as the percent elongation at fracture, i.e., % EL =

Lf L 0 L0

100 ,

(5)

where L0 is the initial length and L f = L0 + L is the final length at fracture. The larger the percent elongation % EL, the larger is the ductility. Note that the percent elongation will somewhat depend on the initial length L0 since most of the plastic deformation is confined to the region of the material in which the material necks, or shrinks, during such plastic deformation (Figure 5). Some materials such as concrete, glass, and brittle plastic show no plastic deformation at all and fracture within their elastic region, therefore being extremely brittle (see again Figure 4). Polymers (such as plastics) also show various stressstrain behaviors, depending on their type. For example rubber is highly elastic, capable of large elongations without plastic deformation. Other polymers may show ductile (large plastic deformations) or brittle (little deformation upon fracture) behavior. In addition the temperature may affect the stressstrain behavior dramatically. For example rubber would become very brittle if cooled with liquid nitrogen.

Experiment 8 8-5

Brittle Extremely brittle Stress Ductile

Strain Figure 4. Stressstrain behavior for brittle and ductile materials loaded to fracture. Necking

Fracture Figure 5. A material test sample first necking and then fracturing under tension.

PROCEDURE DataStudio setup Check if the yellow and black plugs of the rotary motion sensor are connected to the digital channels 1 and 2, respectively. Then check if the economy force sensor is connected to the analog channel A, with the TOP mark on the connector facing straight up. Start the DataStudio software and choose Create Experiment. (If this window doesnt appear, click Setup and choose the proper interface, 750 for USB or 700 for SCSI interface.) If the interface is unavailable, check to see if it is on. If it is on, yet unavailable, and the type of the interface is not USB (but SCSI), you need to restart the computer. In the Experiment Setup window, click Add Sensor and in the pull-down menu select Digital Sensors and choose Rotary Motion Sensor. Then click Add Sensor again and this time choose Force Sensor (not the Student Force Sensor) under Analog Sensors.

Experiment 8 8-6

Click on the rotary motion sensor icon in the experiment-setup window and open the Rotary Motion Sensor tab. Make sure to change the resolution to High. This way the rotary motion has a resolution of 1 /1440 of a full revolution, meaning it can resolve 360 /1440 = 0.25 . Change the Sample Rate to 20 Hz. This way the computer will collect 20 data points every second, or one data point every 0.05 s. In the Experiment menu, go to Set Sampling Options and open the Delayed Start tab. Select Data Measurement. Then select Force, Channel A, and Is Above. Type 0.500 N in the box. This way the data measurement will take place when the condition Force, Channel A, is above 0.500 N is satisfied. In the Set Sampling Options window, open the Automatic Stop tab. Set the condition Data Measurement Force, Ch A is above 50.000 N. This way the data collection will automatically stop when the force on the force sensor exceeds 50 N, which is the maximum allowed for the force sensor.

DataStudio uses variables to analyze Data Measurements. The variables are listed under Data in the left column. For example the rotary motion sensor can measure the Angular Position, which is a variable. Then you can use the Calculator tool to derive other variables from the Angular Position variable (e.g., the angular position multiplied by a constant). Note that the Calculator tool can also be used to define constants that you might need. Finally also in the left column of DataStudio, there are the Display options, which are used to plot these variables. Now it is time to define the constants and variables that we will use in this experiment. Click the Calculate button to open the Calculator window. Click on the Experiment Constants and click New. For the name of the constant, type length. For the value and unit of the experiment constant, type 80 and mm, respectively. Then click Accept. Define another experiment constant by clicking New. Set the name, value, and units to cross_sectional_area, 0.305, and mm^2, respectively. Then click Accept.

Experiment 8 8-7

Define the first variable as follows. Type distance=angle/360 as the formula for the variable in the Definition. Click Accept. For angle choose Angular Position, Channels 1 & 2, under Data Measurement in the pull-down menu. Click Properties and change Variable Name to distance and Units to mm. Click Accept. After you click Accept, click New. Define a new variable by typing stress=force*5/cross_sectional_area. Click Accept. Define force as Force, Ch A. Click Properties and change Variable Name to stress and Units to MPa. Click Accept. Generate a graph of distance vs. stress as follows. Drag the distance icon under Data in the left column onto the Graph icon under Displays at the bottom of the left column. This will generate a distance vs. time graph. Then drag the stress icon under Data on the time axis of the newly generated graph but release the mouse only after the time axis is highlighted. This should generate the distance vs. stress graph. If you released the mouse before the time axis was highlighted, delete the graph and repeat the process. Finally click New again in the Calculator window and define strain=(elongation-apparatus_deformation)/length. Click Accept. Define elongation as distance vs. stress under Data Measurement. Leave apparatus_deformation undefined for now. Also generate a Digits display for Force, Channel A, by dragging it to Digits under Displays. Calibration CAUTION: The maximum force the force sensor can withstand is 50 N. When the force exceeds this maximum, the data collection will automatically stop. Monitor the force constantly on the digits display and stop turning the crank immediately when the force reaches 50 N and the data recording stops. CAUTION: You should turn the crank with light finger pressure only! Never use force when turning the crank. When the crank screw hits one of the stops (either clockwise or counterclockwise), stop turning the crank immediately. If you force the crank screw beyond the stops, the apparatus will be damaged. Figure 6 shows the complete setup for the tensile test. The tension to the test material is applied by turning the crank clockwise. The test material is secured between the clamps on both sides (Part no. 9). Figure 7 is a schematic diagram of these clamps. Note that the serrated surfaces

Experiment 8 8-8

face inward and the thicker clamp is at the top and the thinner at the bottom.

Figure 6. Complete setup.

(a) Regular data run


Bolt Nut Washer Upper clamp (thick) Material sample Lower clamp (thin) Apparatus

(b) Calibration run


Bolt Nut Washer Upper clamp (thick) Calibration bar Apparatus

Figure 7. Positioning of the clamps (schematic). (a) For a regular data run with a material sample and (b) for the calibration run with the calibration bar. Note that the serrated surfaces of the clamps face inward. Also note that the thicker clamp is at the top and the thinner clamp is at the bottom. The tightening is accomplished via a nut and a washer. For the calibration run, only the top clamp is used.

Now remove both nuts using the socket tool, being careful not to loose them or the washers. Also set the four clamps aside. Adjust the positioning of the moving arm by turning the crank and insert the calibration bar onto the two bolts through its holes (Figure 8). After inserting the calibration bar on the apparatus, replace only the top clamp [Figure 7(b)]. Then tighten the nuts using the socket tool. Note: If you are not satisfied with the data run you are just about to make, make sure to first delete it using Delete Data Run under the Experiment menu before you make another data run. This should also apply to any other unsuccessful data run.
Figure 8. Calibration.

Figure 9. Start position.

Experiment 8 8-9

Then make sure that the lever is at the start position (Figure 9) and then tare (zero) the economy force sensor by pressing the tare button on the sensor. After taring the force sensor, click start on DataStudio. Look at the digits display and see if shows zero. Now keep looking at the digits display and, at the same time, very slowly turn the crank clockwise. (You might need the hold the apparatus down with the other hand). As soon as the force on the digits display reaches 50 N, stop turning the crank so that you dont damage the force sensor. At this point the data recording should have automatically stopped. Then relieve the stress on the calibration bar and apparatus by gently turning the crank slightly counterclockwise to bring the apparatus back into the start position (Figure 9), making sure not to force the crank screw after it hits the stop. The calibration run should look very linear (see the figure). Now rename this data run by left-clicking on Run #1 in the Data column for three seconds and then releasing the mouse. Then type calibration instead of Run #1 to rename this data run. To complete the calibration procedure, open the Calculator window and go to the definition of strain (using the pull-down menu next to Definition), if not already there. Now define apparatus_deformation as distance vs. stresscalibration under Data measurement. Make sure to choose only calibration (Run #1), not the entire measurement. DataStudio should give you single data run selected warning, which means that you did the right thing. Click Yes. Then click Accept. Now do you see what the calibration does? The apparatus deforms slightly under stress, since it is not 100% stiff. The calibration bar is very thick and has negligible strain under stress. During the calibration run, we simply measure the strain on the apparatus itself (which is rather small). In the subsequent runs, DataStudio will subtract this strain from the measured strain for the test material, therefore taking the apparatus deformation into account. Dimension measurements Using the digital caliper, measure the width and thickness of each sample and calculate its cross-sectional area. Fill out Table 1 below. Also measure the width and thickness after each tensile test and record them in the table.

Experiment 8 8-10

Table 1. Initial and final dimensions of the sample material.


Initial width mm Aluminum Brass Plastic Polycarbonate Soft steel Hard steel Initial thickness mm Initial crossFinal sectional Final width thickness area mm2 mm mm

Data runaluminum Now we are ready to measure actual stressstrain curves for our test materials. You are supplied with a test coupon each for these six different materials: aluminum, brass, plastic, polycarbonate, soft steel, and hard steel. Note: Do not bend, fold, or crease the test samples; otherwise, they will break prematurely. First remove the calibration bar and assemble the clamps, washers, and nuts on the apparatus exactly as in Figure 7(a) (but dont insert the test material yet). Go to DataStudio and generate a stress vs. strain graph as follows. Drag the stress icon under Data onto the Graph icon under Displays. This will generate a stress vs. time graph. Then drag the strain icon onto the time axis of the newly generated graph but release the mouse only after the time axis is highlighted. If you make a mistake, delete the graph and repeat the process. (See the figure.) Turn the crank all the way counterclockwise so that the moving arm is all the way in. Now double-check if your clamps conform to Figure 7(a) and then insert the aluminum test sample between the clamps as in Figure 7(a). Do these for the clamps on both sides. Do not tighten the nuts yet. At this point make sure that both ends of the test sample are completely pushed against the bolts, and the test sample is not twisted diagonally (see Figure 10). Once the test sample is correctly positioned as in Figure 10, tighten the nuts using the socket tool. Make them

Figure 10. Correct placement of the material test sample.

Experiment 8 8-11

tight so that the material doesnt slip, but do not use any other tool, which may break the bolts and damage the apparatus. Also make sure that the lever is at the start position as in Figures 9 and 10. Now first press the tare button and then click Start on DataStudio. Turn the crank very slowly clockwise and watch the digits display. Stop turning the crank immediately if the force reaches 50 N. If the sample breaks, stop turning the crank immediately as well and manually stop DataStudio by clicking the Stop button. After the data run is complete, rename this run (i.e., Run #2) aluminum as you did when you renamed Run #1 calibration. CAUTION: In DataStudio it is sometimes necessary to click Accept in Experiment Constants in the Calculator window again, if your graph gives a division-by-zero error (i.e., number out of range) or has some unknown problem. Other data runs Loosen (but do not remove) the two nuts and remove the aluminum sample. Turn the crank all the way counterclockwise and then place the brass sample. Tighten the nuts and make sure that the apparatus is in the start position. Tare the force sensor and start DataStudio. Very slowly turn the crank clockwise and let the program take the data. Constantly monitor the force on the digits display at the same time so that you dont exceed 50 N. When done, stop DataStudio and rename the data run (i.e., Run #3) brass. Repeat the same procedure for plastic, polycarbonate, soft steel, and hard steel. For hard and soft steel, you need to make the nuts tight enough so that they dont slip between the clamps. Turn the socket tool with your hands to tighten the nuts. Do not use any other tool. CAUTION: With hard steel be extremely careful not to exceed 50 N since this material is very strong and will take large forces. ANALYSIS Now we will analyze our results. When you fill out Table 2 below, for each material make sure to first go to the Calculator window in DataStudio, and then enter the correct initial crosssectional area (in mm2) for that material from Table 1 and click Accept. This way DataStudio will use the correct initial cross-sectional area in the calculations for each material. Also in the stressstrain graph, select only the data run you are analyzing by using the Data icon .

Experiment 8 8-12

First calculate the Youngs modulus E as follows. Make a linear fit using the Fit icon. Using the mouse select the linear region of the curve but ignore the nonlinearity at the beginning, which is caused by the fact that we use a two-dimensional sample, which is not stiff. The resulting slope is the young modulus, in this case 24,300 MPa or 24.3 GPa (gigapascal). (See the figure on the right.) Since aluminum doesnt show a yield-point phenomenon, such as in Figure 3, we need to use the method in Figure 1 to find the yield strength y . This time make a User-Defined Fit and select Manual in the Curve Fit window that comes up. For the equation of the curve, enter 24300*(x-0.002), where 24300 is the slope and 0.002 is the x-intercept. Click Accept. (See the middle figure on the right.) Then read the yield strength y from the intersection of this line with the stressstrain curve. You can also use the

xy tool to find the intersection. For aluminum in the example here, the yield strength y is 130 MPa. (See the bottom figure on the right.)
If the material that you are showing exhibits a yield-point phenomenon, then the yield strength y is measured in a way similar to in Figure 3. Therefore use the method in Figure 3 for such materials, instead of drawing a parallel line. Note that for plastics, the upper yield point, instead of the lower yield point as in metals such as certain types of steel, is taken as the yield strength y . The tensile strength TS is just the maximum stress on the stressstrain curve. The breaking stress is the stress at the fracture. The ductility is measured by the percent elongation % EL, which is the amount of elongation at fracture, given by % EL = 100 ( L f L0 ) / L0 = 100 f , where L0 and L f are initial and final

lengths and f is the strain at fracture. Therefore for aluminum in the example,

% EL = 100 0.010 = 1.00% . A material is considered brittle if the fracture strain (percent elongation % EL at fracture) is less than about 5% and is considered ductile otherwise.

Experiment 8 8-13 Table 2. Analysis.


E

TS
(MPa)

(GPa) Aluminum Brass Plastic Polycarbonate Soft steel Hard steel

(MPa)

Breaking stress (MPa)

% EL %

Fractured

Yield-point Necking Ductile or phenomenon phenomenon brittle Yes/No Yes/No Yes/No Ductile/brittle

At this point discard all the used material samples into the trash.
Final activitybeam deflection

When we studied the tensile tests for various materials in this experiment, we learned how different materials reacted to forces that put them under tension. In this activity we will study the deflection of beams with different cross-sectional shapes built out of the same material. You are supplied with three different beams made out of the same kind of steel with the following cross-sectional shapes: H, L, and T. All beams have approximately the same mass density per unit length. The beams will be supported by two metal supports at both ends. An indicator gauge will be placed under the beam in the middle to measure the deflection under load.
Table 3. Beam deflection.

Place the beam on the stands and place the indicator gauge in the middle, its height adjusted so that the indicator needle is slightly pressed by the bottom surface of the beam. Make sure that the body of the indicator gauge doesnt interfere with the beam. Then zero the indicator gauge. Stand on the beam in the middle and have your partner read the indicator gauge. Record the reading in Table 3 below. Repeat the same procedure for the other two kinds of beams.

Beam deflection in the middle mm H beam L beam T beam

Question 1. What seem(s) to be the geometrical factor(s) contributing to the strength of a beam?

Experiment 8 8-14

Name: Partners name: Date and time:

PHYS135AL

EXPERIMENT IX FLUID FLOW

1.0 Objective
In this experiment we will be examining a property of liquids called viscosity. A series of experiments will be performed to verify the validity of mathematical relationships regarding fluid flow and to determine the viscosity of water.

2.0 Background
All liquids display a property called viscosity - a measure of a fluids thickness (tendency to resist flow.) In essence, a fluids viscosity is determined by the internal frictional force between different layers of the fluid as they move past one another. In liquids, this friction is due to cohesive forces between molecules. All fluids possess different levels of viscosity, though often viscosity is not specifically appreciated as a characteristic property of a liquid (water in daily usage). Yet on occasion it becomes very apparent (maple syrup). Viscosity of liquids is temperature dependent. 2.0.1 What happens to maple syrup when you heat it? when you cool it?

The viscosity of fluids can be expressed quantitatively by the coefficient of viscosity, . To better define , consider the following: A thin layer of fluid is placed between two flat plates. One plate is stationary, the other is made to move (figure 1). The fluid directly in contact with either plate is held in place by cohesive forces between the molecules of the liquid and those of the plate. Thus the upper surface of the fluid moves with the same speed, v, as the upper plate and the liquid in contact with the stationary plate remains stationary.

EXP IX

Moving Plate F v . Statio nary plate Figure 1 Laminar flow Between each thin layer of fluid, friction retards the flow of the layer just above it and so on through the entire liquid. The result is that the velocity of the fluid varies linearly from 0 to v as shown in figure 1. To move the upper plate in this example requires a force. For a given fluid, this force, F, is proportional to the area of either plate, A, and to the speed, v, and is inversely proportional to the separation, L, of the plates: F v A / L. The proportionality constant is defined as the coefficient of viscosity, L

F = Av --------L

(EQ 1)

2 2 The units of are Ns / m , which is equivalent to Pas (Pascal is the standard unit of pressure, defined as 1 N/m ).

2.1 Categorizing flow


In general, we can distinguish two types of fluid flow: laminar flow and turbulent flow. Laminar flow (or flow in layers) describes a regular flow pattern as depicted in figure 1 and used to derive equation (1). All aspects of laminar flow can be fully characterized both physically and mathematically. Turbulent flow describes all remaining forms of fluid flow, characterized by irregular flow patterns like vortices, intersecting stream patterns, etc. For this type of flow, simple physical description, prediction or mathematical characterization is no longer possible. In fact, advanced chaos theory is often used to obtain some form of description and understanding of turbulent flow. The onset of turbulent flow is often abrupt. For a cylindrical tube, turbulent flow can be characterized approximately by the Reynolds number, Re:

Re = 2 < v > R -----------------------

(EQ 2)

where <v> denotes the average speed of the fluid, the density of the fluid, R the internal radius of the tube, and the coefficient of viscosity of the liquid. Flow in a cylindrical tube is laminar if Re < 2000, but is turbulent if Re 2000

EXP IX

2.1.1 The aorta has an approximate radius of R 1.0 cm. The average speed of the blood during the resting cycle is about 30 cm/s. During this cycle, is the flow in the aorta laminar or turbulent? (For blood the density is 1.05 103 kg / m3 and the viscosity 4.0 10-3 Pas)

From this point on, we will be examining aspects of laminar flow only.

2.2 Laminar flow


If a fluid had no viscosity, it could flow through a level pipe without any force being applied. Because of viscosity, a pressure difference between the ends of a tube is necessary for steady flow of any fluid, be it water, maple syrup or blood in the circulatory system of a human being. The volume rate of fluid flow in a cylindrical tube, Q, is defined as the volume of fluid flowing past a given point per unit time. Q depends on the viscosity, the pressure difference, and the dimensions of the tube according to the Poiseuille equation:

R P Q = ----------------8L

(EQ 3)

In EQ 3, R denotes the inside radius of the tube, P the pressure difference (also called drop) between the ends of the tube, the viscosity, L the length of the tube.

3.0 Procedure
We will now progress into a set of experiments allowing the opportunity to determine the viscosity of water at room temperature. You will need the following equipment: A 5 gallon water reservoir, feeding liquid-distributor box, which, in turn, feeds a selection of tubing of various lengths and diameters;a set of containers to catch water flowing from the tubing; a balance, allowing accurate determination of the mass of the collected water; a stop watch, for accurate registration of time intervals.

EXP IX

Figure 2: experimental set-up


A

Pipe Rise tube

Outflow opening
B

Liquid distributor box Tubing connectors


When performing the flow experiments, the water level in the supply reservoir will gradually drop. Since the elevation of this water level strictly defines the pressure of the flow system, we should find that during a measurement sequence the pressure will gradually drop with the decrease in the reservoir. Our system actually overcomes this problem as follows: Consider our supply reservoir in figure 2, which has no other openings than an outflow opening B and a hollow pipe A. When water is drained from the reservoir through B, air is pulled in through A to restore the pressure balance between the inside of the reservoir and outside atmospheric pressure. As a result, when the reservoir is drained, the pipe at A will remain free of water and air bubbles as air enters the reservoir. Consequently, the pressure inside the reservoir at the lower end of the intake pipe is always identical to atmospheric pressure. The pressure in the liquid-distributor box is measured by a rise tube. The static pressure in the box will cause the water column to rise in the tube. The height of the column is proportional to the pressure inside the box and can be read on a ruler. The gauge pressure inside the distributor box is then calculated using the formula:

P = gh

(EQ 4)

where = 1.0 x 103 kg/m3 is the density of water, g the acceleration of gravity, and h the height of the water column. Flow and pressure are controlled by opening and closing the reservoir spigot by means of a hose clamp.

EXP IX

3.1 Determination of the viscosity of water


Measurements on the flow system will generally use equation (3): Q, the output of the flow system will be measured as function of R, the radius of the tubing, P, the pressure difference over the tubing, and L the length of the tubing. As stated earlier, the coefficient of viscosity is temperature-dependent. For this measurement series use the tubes with the smaller diameter of 1/32". 3.1.1 Note the temperature in the room as an important factor influencing viscosity. T =?

The flow rate, Q [m3/s], is obtained by measuring the outflow of water in grams for a fixed period of time. 3.1.2 How would you calculate the volume of water outflow from the mass? (Hint: use the density of water).

3.1.3 How would you use this result to then calculate Q?

First establish equilibrium pressure in the setup: The rubber stopper on the cap of the water tank should be tight. Open the tap of the water tank. Drain water from the setup until the water level in the rise tube stabilizes. The height of the water in the rise tube, measured from the level of the cocks, corresponds to the pressure difference P across the small vinyl tubings, since one end of the rise tube as well as one end of each tubing are open to the air. Leave the tap of the water tank open for the entire experiment. Note the room temperature in your answer book. 3.1.4 Measure the height of the water column and use it to determine the pressure difference over the tubing from EQ 4. Note: The zero point of this measurement should be the height of the distribution connectors.

EXP IX

Now you are ready to let the water run through the tubes. During the experiment: make sure, that the outflow level of each tube is on the same height as the inflow level! NOTE: If the tube seems to be blocked, use the injector to remove the air and push the water. 3.1.5 Measure the outflow for tubings with an internal diameter of 1/32 " (1" = 2.54 cm) and various lengths (50 cm, 33 cm and 25 cm). Measure the outflow for at least 15 minutes for each length. NOTE: Taking turns with your partner, outflow measurements on the different lengths of tubing can be performed virtually simultaneously, substantially reducing experimentation times.

Length (m)

mass (kg)

Vol. (m3)

t (sec)

Q (m3/s)

3.1.6 From the outflow measured, calculate the output, Q, and express it in m3/ s.

EXP IX

3.1.7 From your values make a graph Q vs. 1/L on the graph paper provided.

3.1.8 Determine the slope for an eyeball-fit line for your graph (assume the line passes through the origin.)

3.1.9 Determine the viscosity of water, using your value for the slope, the radius of the tubes and the pressure difference along the tubing from EQ 3.

EXP IX

3.1.10 How would the flow rates change, if all the used tubes had twice the diameter than the ones you used? Would the value for the viscosity of water change?

3.1.11 How would the flow rates change, if the height difference between the reservoir and the distributor box is increased by a factor of two?

3.1.12 Explain how the flow rate would change if in the original set-up oil is used in place of water.

3.2 Dependence of laminar fluid flow on the shape of the flow path
According to equation (3), Q is only dependent on the pressure across the length of tubing, i.e., the pressure difference between both ends of the tubing. Thus Q is independent of the path taken by the fluid. We will verify this, and relate this phenomenon to blood flow in the human body. For this experiment you use the tubes with the larger diameter. Using longer lengths of tubing, call upon your creative self to construct the four different shape pathways listed below for the flow to travel (be certain to measure Q in all situations) [Be careful not to kink, crush, knot or tightly coil the tubing]: 1. curving upward, then down (mountain); 2. down, then up (valley); 3. sideways, then back (moderate curvature); 4. moderately coiled.

EXP IX

3.2.1 Record relevant data and calculate values of flow for each of the four cases.

3.2.2 What can be said about the values of Q obtained in situations 1, 2, 3 and 4?

3.2.3 How would you translate the importance of this into blood flow related aspects in the human body?

3.2.4 List sources of error for both experiments and state how each error affects Q.

EXP IX

EXP IX

10

Name: Partners name: Date and time:

PHYS135AL

EXPERIMENT X SOUND

1.0 Objective
To study the characteristics of sound waves and combinations of notes.

2.0 Background
2.1 Sound as a wave
Wave motion is a phenomenon seen in many aspects of everyday life, from ripples in a rain puddle to the roar of a jet flying overhead to radio and television programming to the evenings violin concerto. Mechanical waves, waves which are generated in any elastic medium by rapidly displacing a portion of that medium, are the most familiar variety. Mechanical waves transmit energy and momentum from point A to B without actually causing any of the medium in which they travel (propagate) to relocate from A to B. Some local disturbance (at point A) initiates a local vibration (oscillation) in the medium which is either parallel to (longitudinal [or, compression] waves) or perpendicular to (transverse and torsion waves) the direction the wave disturbance actually travels. Some time later the disturbance in the medium is detected at a different location in the medium (point B) having been transmitted by action and reaction within the medium rather than its actual, permanent displacement. (Think of The Slinky, which exhibits compression, transverse, and torsional wave motion, depending upon how you start it vibrating.) 2.1.1 Give an everyday example of a longitudinal wave. Draw a snapshot of the wave that shows how the medium is displaced. Then being careful not to permanently bend, twist, or deform Slinky, place it on the table with your partner holding one end while you gently generate a longitudinal wave along it. Note that the direction of wave propagation is from you to your partner.

2.1.2 Give an everyday example of a transverse wave. Draw a snapshot of the wave that shows how the medium is displaced. Then generate a transverse wave along Slinky as you did above.

EXP X

2.1.3 Generate a torsion wave along Slinky. Describe how the torsion wave is similar and different compared to a transverse wave. Then draw a snapshot of the wave that shows how the medium is displaced. Can you give an everyday example?

2.1.4 What does the working definition of mechanical waves above imply about propagation when no medium is present (i.e., in vacuum)? Explain.

Sound is a mechanical wave, more specifically it is a longitudinal mechanical wave. It can actually be caused by all sorts of other wave types, including transverse waves as produced by stringed instruments. A disturbance originates at point A (original cause can be anything), and then compresses the medium (usually air) which then reacts to that compression. The mediums chain reaction allows the disturbance to propagate at a velocity the frequency f , and wavelength

,which depends upon


(EQ 1)

, through the medium until dissipative forces (friction, heating) damp it out. = f

Both

and

depend upon the medium; the intensity of a sound depends upon the medium; the frequency depends

only upon the source. The speed of sound in air at 20 0 C is 344 m/s. 2.1.5 Would you expect a specific sound to sound different if heard under water? Explain in detail.

As the disturbance passes your ear, the oscillating medium in turn causes an oscillation in your ear which you then perceive as sound IF it is within your hearing range -- i.e., if its frequency falls within the range to which your ear is

EXP X

sensitive -- and IF its intense enough to actually stimulate your hearing mechanism when it arrives. Clearly then our sense of hearing depends upon frequency and intensity [W/m2] (which is directly related to POWER = energy per unit time [J/s = 1 Watt]) considerations. The threshold of human hearing is at an intensity of approximately I0= 1x10-12 W/m2 and ranges up to 1 W/m2 or greater. The 12 powers of ten (or more) in intensity range makes a logarithmic scale more convenient. Thus audible intensity ranges are defined in terms of I0 and decibels, where the intensity level of a sound with intensity I in decibels is given by:

I = 10 log ---I0
The typical frequency range of the human ear is also large, ranging from about 100 Hz to circa 15 kHz.

(EQ 2)

2.1.6 A given sound has an intensity of I. Now the source moves 2 times further away, meaning the intensity decreases by a factor of (2)2 = 4. What is the audible change in the intensity level (in decibels)? (HINT: use log(x / y) = log(x) - log(y) to calculate the difference in decibels.)

2.2 Elementary music theory


Musical notes are sound waves with pure frequencies. Middle A has a frequency of exactly 440 Hz. Middle C has a frequency of approximately 262 Hz and so on. A note which has a frequency of exactly twice that of another is said to be an octave higher, it is also called the first harmonic of the original note, which is then called the fundamental. The note which is one octave above the first harmonic is called the second harmonic, and so on. You might think, that when a musician plays middle A his instrument produces a perfect wave with a frequency of 440 Hz. In fact, this is almost never the case (electronic instruments, such as the tone generator on your computer do produce pure frequencies). When a musician executes a note, higher harmonics are excited in the instrument as well, but with lower intensity. It is the combination of the fundamental and its associated harmonics that differentiates the sound quality we perceive both between instrument types (is it a guitar or a violin?), and instruments of the same variety (if its a violin, is it a good violin?). The specific combination of the fundamental and its associated harmonics acts as a sound fingerprint which identifies the source. These fingerprints are better known as the sound spectrum of an instrument or source. For an instrument, the sound spectrum is also called the harmonic content. Every wave source has a spectrum, quite often represented by a graph of intensity vs. frequency for the source in question.

EXP X

One way of combining notes is to play a chord. This is done by playing together different notes, the frequencies of which are related by simple fractions. For example, if one plays the chord C major, one plays the notes C, E and G together. The ratios of the frequencies are in this case:

fE fG ---- = 5, ---- = 3 -- - -fC 4 fC 2

For different types of chords, the frequency ratios will be different. If we play notes together which have a frequency ratio that is not a simple fraction, the resulting sound is not pleasant to human ears.

Figure 1: frequency rages of various voices and instruments

EXP X

2.3 Combining waves, and beats


After learning about music theory, consider the mathematical expression for two pure sound waves (notes) of amplitudes A and

B, and consisting of single frequencies f 1, 2 :

y 1 ( t ) = A sin ( 2f 1 t ) y 2 ( t ) = B sin ( 2f 2 t )

(EQ 3) (EQ 4)

Waves add to one another using the superposition principle - the linear combination of two or more waves. Then a sound consisting of these two waves would have the form of the simple mathematical addition of EQ 3 and 4. 2.3.1 Assuming identical amplitudes, add the two waves in EQ 3 and 4.

2.3.2 Now use the mathematical identity for sin addition sin ( x ) + sin ( y ) = 2 sin ----------- cos ---------- to simplify the

x+y 2

xy 2

equation.

We now define

f1 + f2 ------------- f = the average frequency of the combined waves 2


and

(EQ 5) (EQ 6)

f 1 f 2 f' = the beat frequency.

EXP X

2.3.3 Rewrite your solution to 2.3.2 in terms of average frequency and beat frequency.

In general, when you add two waves with distinct frequencies, you get a single wave with two characteristic frequencies. Beat frequencies are detectable - thus the importance of having all the instruments in an orchestral section exactly in tune with one another becomes very clear.

3.0 Procedure
We will now graphically explore notes, beats and other aspects of sound using a PC, a signal-processing program called Cool Edit, a computer microphone, and two tuning forks.

3.1 Cool Edit set-up


You will be using the computers built-in microphone, which should already be properly connected. Step 1: Power up your PC and double-click on the 135a folder. Your computer will already have an external microphone installed, or come with one built-in, and be ready to use. Double-click on the Cool Edit icon to start the program. Step 2: From the menu bar above, select file: new. The window in Figure 2 will pop up. You will need to set the following properties: SAMPLE RATE = 22050. CHANNELS = STEREO. RESOLUTION = 16-bit. Click ok to continue.

Figure 2: set-up window

Step 3: A new session titled Untitled with top and bottom panels will come up. The top panel is for the left audio track of the stereo input and the bottom track is for the right channel of the stereo input. To select one of the channels, click on the top-left or bottom-left corner of the panel. Depending upon which channel you choose, the cursor will change from an arrow with a L below it or a R below it. You are now ready to use Cool Edit to examine sound samples.

EXP X

HELPFUL HINTS: Hint 1: When you record a sound, make certain it doesnt saturate the recording device. You know your recording is too loud when the recorded waveform disappears above and below its window. Experiment with different input audio levels by moving the source away from the microphone until the wave is no longer clipped, but still audible. Hint 2: Save your sound files at regular intervals to My Music, choosing a unique filename. Recycle this filename as you go through the various sections to save disk space.

3.2 Sound wave characteristics


Lets look at a fundamental sound wave and its associated harmonics. In the Cool Edit menu, go to File: Open, and open the fundamental sound file in the 135a/sounds folder. A new window with this file will appear on your screen and the untitled window will disappear (but it is still there, open). Play with the window controls on the fundamental file until you feel comfortable with them. Click on the fundamental window. Select the entire waveform. [Click and drag along the wave, being certain to highlight the entire wave]. In the menu bar, go to Edit and choose Copy. In the menu, go to File: Close to close fundamental, making certain NOT to save any changes. The two-track untitled window should reappear. Click on the top channel (left track) at least once to select that window. In the menu bar chose Edit: Paste. The fundamental wave form will be pasted into the top untitled window panels. Adjust the amplitude if needed by going to the menu and choosing Transform: Amplitude: Amplify. Now open the file called 3rd harmonic. A new sound window will pop up with this file. Using the method above, copy 3rd harmonic into the bottom (right channel) empty panel in untitled. Go to File: Close to close 3rd harmonic, without saving changes. Note that both the fundamental and the 3rd harmonic are pure (single) frequency waves. 3.2.1 By inspection, determine the ratio of the frequency of the 3rd harmonic to the fundamental. You should not use the Amplify function but keep the amplitude setting at 100%.

In the menu, select Edit: Mixpaste. Save this new wave as sum-one.

Figure 3: settings for Mix Paste panel

EXP X

3.2.2 Sketch the new waveform.

Go to File: Open and select 5th harmonic. Using the exact same steps, add a track and mix the 5th harmonic to sumone. Save this wave as sum-two. 3.2.3 Sketch sum-two. Repeat the process until youve added the 7th and 9th harmonics to each successive summed wave. Sketch each intermediate sum.

3.2.4 Qualitatively speaking, what is/are the general trends of the summed wave as each successive harmonic is added to the previous?

3.2.5 Draw a general conclusion about the net sound wave of an instrument which has a high harmonic content.

EXP X

3.2.6 Name a relatively strong fundamental instrument. Name a relatively high harmonic instrument.

Generally speaking, waves which have a jagged shape sound more harsh to the ear than instruments that have stronger fundamental components. Make certain your lab instructor checks your final summed wave before you go to File: Close to close all these windows and move onto the next section.

3.3 Notes and chords


Cool Edit provides us with an electronic instrument capable of easily producing pure frequency waves. In the menu, first go to File: New and open a new stereo panel. Now choose Generate: Tones. Select the note you want to hear start with a 400 Hz tone.

Figure 4: settings for Generate Tones panel

Adjust the dB volume to its highest value and click ok to see the wave form. Slide the wave around until youre convinced it is indeed a sine wave. Create tones of 500 Hz and 600 Hz. Play each of these notes so you get a feel for each individual tone.

EXP X

3.3.1 Determine the ratios of the last two frequencies to the first one. Explain the actual process you use to do this in detail.

3.3.2 Do these ratios agree with the values given in Section 2.2? Do you expect to like the sound, when you play the mixed wave?

Using the instructions in 3.2, mix the three tracks until you have one final wave which is the sum of all three notes. But this time normalize the amplitude of each wave to 30%. Play the result. You should hear a major chord. 3.3.3 Sketch the final wave form here. Can you see the influence of all three frequencies in this chord?

Make certain your Lab Instructor checks your final summed wave before you close these windows and move onto the next section.

EXP X

10

3.4 Beat frequencies


Open up a a sine wave of 265 Hz and a sine wave of 267 Hz (both with an amplitude of 50% and duration of 5 sec). 3.4.1 Play back both notes one after the other - do you hear any difference?

Now mix the two waves and play back the resulting wave. 3.4.2 Describe what you hear now.

What you hear is precisely the beat, modulating the amplitude of the average sound wave. 3.4.3 For this experiment, calculate the frequency of the beat and the average sound wave.

3.4.4 Repeat all of the above steps by now combining 265 Hz and 275 Hz and sketch the mixed wave.

11

EXP X

In your equipment box, you should have two aluminum tuning forks. Were going to record the output from each tuning fork using the computers microphone, then examine the two characteristic frequencies formed when these two recorded waves are added (mixed). Go to Menu: New and open a new, empty sound window. Click on the window, then holding the first tuning fork by its stem, carefully strike one of the prongs with the rubber mallet. [Do NOT strike the tuning forks on any surface, or hit them with anything besides the mallets.] The vibration of the prongs produces a pure frequency sound, the numerical value of the frequency is stamped on each fork. Hold the tuning fork STEADY a few inches from the microphone and click Record in the bottom left corner of the screen. Record a ten-second tone for this fork. You need to maintain as constant amplitude for this wave as possible. The amplitude is controlled by the forks proximity to the microphone. You might need to redo this recording a couple of times before you get a reasonably constant amplitude. Once you have a good recording use Transform: Amplitude: Amplify and use the Normalization feature to set the amplitude to 40%. (Click Calculate Now and then click OK.) Then go to the Edit menu and Copy the wave. Now record the second tuning fork for 10 seconds. Once you are done, set the amplitude of this wave to 40% as well. You are now ready to add the two waves together. Use the Mixpaste in the Edit menu to do this. 3.4.5 Sketch the mixed wave.

f and f' which are

Notice that the mixed wave has a fast average frequency of oscillation a slower frequency

f . Superimposed over the entire wave form is

f' . Play back the mixed wave and see if you can hear BOTH of these frequencies.

3.4.6 Describe what you hear.

EXP X

12

3.4.7 From the graph, estimate both the average frequency and beat frequency. You should first set the time-axis units to hms (hours, minutes, seconds) by double-clicking on it. Remember that the period of oscillation is the inverse of the frequency (i.e. T = 1/f). Indicate in your sketch which periods you used to determine these two frequencies.

3.4.8 Now use EQ 5 and 6 to calculate the theoretical values of f and f' , using the frequencies stamped on the tuning forks. Compare the results with your estimates from the graph.

Before you leave the lab today, try holding one of the tuning forks in front of the computer screen - you will then not only hear but also see the vibration of the fork! Can you explain this?

13

EXP X

EXP X

14

Name: Partners name: Date and time:

PHYS135AL

EXPERIMENT XI IDEAL GAS AND THE ABSOLUTE ZERO OF TEMPERATURE

1.0 Objective
To study the variation with temperature of the absolute pressure of a dilute gas held at constant volume. To determine the absolute zero of temperature via extrapolation of your data.

2.0 Background
In 1662 Boyle found that for a fixed amount of gas maintained at a constant temperature T, the product of absolute pressure P and volume V was a constant. The relation is known as Boyles Law:

PV = constant

(EQ 1)

A century later Charles and Gay-Lussac found similar relationships for pressure and temperature, and volume and temperature. All three results are consequences of the ideal-gas law:

P abs V = nRT K
where: Pabs = absolute pressure V = volume of gas

(EQ 2)

TK = absolute temperature = temperature in the Kelvin scale n = number of moles For this experiment: R = the universal gas constant = 8.314 Joule/(mole K)

P abs = P atm + P gauge


The gauge pressure

(EQ 3)

P gauge is the difference of the absolute pressure P abs from the atmospheric pressure

P atm . It is a useful quantity since most pressure gauges, such as the tire gauges you use to check the air pressure
in your tires (28 psi, 32 psi, etc.), measure the gauge pressure, i.e., the difference from the atmospheric (open-air) pressure. If we seal a fixed quantity of gas in a container with volume V constant, then EQ 2 becomes:

P = nR T P = constant ------ T V

(EQ 4)

EXP XI

Here T is typically measured by a thermal expansion thermometer such as a mercury thermometer. Two liquid thermometers using different substances will not always give the same reading for the same temperature, even if both have been carefully calibrated. This occurs because the volume expansion coefficients of liquids are not independent of temperature. However, we can use dilute gases to give us a universal definition of temperature as they all return the same result within the accuracy of our experiment.

2.1 Temperature
A quick diversion into the subject of temperature: Temperature scales are all relative, depending upon one definition as a reference point. The Fahrenheit scale defines the freezing point of water to be 32o, and the incremental scale changes such that the boiling point of water occurs at 212o F. By comparison, the Celsius scale defines the freezing point of water to be 0o C, using a different incremental scale to set waters boiling point equal to 100o C. Either scale is both useful and valid; scientists typically use the Celsius scale for a variety of reasons, but conversion between the two is easy:

9 o T F = -- T c + 32 5 5 o T c = -- [ T F 32 ] 9
One must always be careful to specify which you are using when reporting data.

(EQ 5)

(EQ 6)

This experiment is interested in finding the absolute zero of temperature- defined as the temperature where the pressure of an ideal gas would become zero. In practice this value approaches a very large negative number in BOTH the Fahrenheit and Celsius scales. So we set up yet another temperature scale, one that has its zero point precisely at absolute zero. Known as the Kelvin scale, it is defined with respect to the Celsius scale as follows:

T K = T c + 273.15
Note: Temperatures in Kelvin are always written without the degree mark (o). 2.1.1 What is the freezing point of water in degrees Kelvin?

(EQ 7)

2.1.2 The boiling point of water?

EXP XI

3.0 Procedure
CAUTION: Liquid nitrogen is very cold and if spilled can cause severe burns and skin damage. Be especially careful to prevent liquid nitrogen splashing onto your face or into your clothes where it can get trapped against your skin. CAUTION: Never put the digital thermometer in liquid nitrogen, which will destroy it. In this lab you will be given two containers with different gasses inside, one of them marked with a green label. The following procedure is the same for each of the gases. We will measure the pressure of each gas at three points: liquid nitrogen boiling, water-ice equilibrium, and watersteam equilibrium (boiling water). Each of the containers has a fixed amount of a dilute gas in it, held at constant volume. You will measure the absolute pressure P of each gas at different temperatures T, plot the pressure as a function of temperature, and analyze your results.

Figure 1: Pressure vessel and temperature baths


Pressure Gauge

Pressure Vessel

Clamp Stand

Hot Water Bath

Water-Ice Bath

Liquid Nitrogen Bath

Each gas is contained in a spherical stainless steel pressure vessel at a room temperature, with a gauge pressure of about 90 PSI (pounds per square inch). The pressure vessel is attached to a handle via a vertical rod. This allows you to move the vessel from one temperature bath to the next with a minimum of effort. You can clamp the vessel at the appropriate height using the clamp stand. Always make sure that the pressure vessel is not in contact with the container walls of the temperature bath. For doing measurements with liquid nitrogen, it is easier if you remove the rod from the clamps and hold the rod with your hand near the handle and gauge. Then insert it into the bath manually and hold it until it reaches equilibrium. But do not

EXP XI

touch the bottom of the rod or the sphere since they will be too cold and may cause burns. After you are done, carefully set it aside, on the cloth provided. The gauge pressure of the gas pressure is measured in PSI (pounds per square inch) by a gauge mounted on the top of the pressure vessel. When you come to class, you will need to begin boiling water to create a 100oC temperature bath. Before the lecture, fill the metal container with hot water so that the water will completely cover the spherical part of the pressure vessel. Turn the hot plate to maximum. It takes about 30 minutes for the water to begin boiling. Remember, do not let the spherical vessel have contact with the metal bath container. When the water starts to boil, reduce the heat to maintain a gentle boil. Be sure that the water level remains above the top of the vessel. 3.0.1 Approximately two minutes after the water has started to boil, record the time and gauge reading. Repeat this at least 4 more times at two to three minute intervals or until three successive readings are identical.

3.0.2 What is the temperature of the gas in the pressure container in oC?

Gently tap the face of the pressure gauge to insure that the needle is not stuck. When you are satisfied that you have accurate readings, turn off the hot plate and remove the spherical vessel from the hot water. Be careful; hot water can cause severe burns. You can leave the hot water for the next lab section. Now move the pressure vessel to the water-ice bath. This bath should be a mixture (50% each) of ice and water. Again, make sure the pressure vessel is not in contact with the sides of the bath container and that the entire spherical portion of the vessel is covered with the bath. Monitor the pressure, after it has remained constant for several minutes repeat the procedure above by recording the pressure and time. 3.0.3 The gas is now at the water-ice equilibrium temperature of T = ______ oC and pressure P = ______psi. Again, you can leave the water-ice bath in place for the next lab section. IMPORTANT: Before proceeding to the next bath, wipe the pressure vessels dry using paper towels. Now fill the blue insulated bath to about half full of liquid nitrogen. Your Lab Instructor will assist you with finding and pouring the nitrogen. Slowly insert the pressure vessel into the insulated nitrogen bath. Again, remember to keep the pressure vessel from contacting the sides of the bath. As the liquid nitrogen boils away, you may need to add more to maintain the liquid level such that it completely covers the spherical part of the pressure vessel. As the pressure vessel approaches the temperature of the liquid nitrogen, the rapid boiling will cease and the liquid nitrogen will exhibit a gentle churning motion.

EXP XI

3.0.4 About two minutes after this is observed, record the time and gauge pressure.

Repeat this measurement at least 5 more times at two to three minute intervals or until the three successive pressure readings are identical. The gas is now at liquid nitrogen boiling temperature T = 77.35 K= -195.80 oC. Remove the pressure vessel from the nitrogen bath and set it aside to warm to room temperature. DO NOT touch it; it is still very cold. You are now ready to finish your calculations. 3.0.5 Record the atmospheric pressure and convert to PSI units. (1 atm = 14.70 PSI = 760 mm Hg = 29.92 inch Hg)

3.0.6 Convert all pressure measurements to absolute pressure using EQ 3, and construct a table of your experimental results. (What should you include?) Estimate uncertainty for each measurement.

Table 1: Experimental data T Gas 1 (C) P gauge


Gas 1

P abs

Gas 1

T Gas 2 (C)

P gauge

Gas 2

P abs

Gas 2

Error in temperature:

EXP XI

Error in pressure:

3.0.7 Plot absolute pressure in PSI vs. temperature in oC for each gas (on the same graph), marking the points for each gas differently. (Note: Make a landscape-view graph, set the pressure as the y-variable and the temperature as the x-variable, and use a temperature range from -400 oC to 100 oC.) Carefully, study your graphs. Can you say that any of your graphs is (approximately) linear?

3.0.8 What kind of graph P vs. T when V=const would you expect for an ideal gas?

3.0.9 How do the graphs for the gases you studied compare to the one for the ideal gas?

3.0.10 What conclusion can you make about the gases you studied today?

Now, let us concentrate on the linear graph for one of your gases. 3.0.11 Use a straight edge to draw the best linear fit to your data. 3.0.12 Extend the line to zero pressure and record the corresponding temperature.

3.0.13 Compare your value with the accepted value of -273.15 oC. What is the percent error of your value?

3.0.14 Now calculate the least square fit to your data and obtain the slope m and the intercept b. Be certain you have three significant figures in your calculations. In order to do this use DataStudio. (Create Experiment, New Empty Data Table from the Experiment menu, drag Editable Data to Graph, Linear Fit. Also on the graph, right-click, go to Settings, and uncheck Connect Data Points.)

3.0.15 Obtain the temperature at P=0 from m and b. Compare this value with your first value.

EXP XI

EXP XI

EXP XI

Name: Partners name: Date and time:

Physics 135AL

Experiment 12 Thermal conduction and radiation: case study and Leslies cube OBJECTIVE To quantitatively investigate the heat flow through the walls of a model structure. To investigate thermal radiation using a Leslies cube. EQUIPMENT One in the room: model structure with four different walls and roof, two infrared heat lamps (GE 37771, 250 W) inside the model structure, and glass thermometer inside the model structure; one for each group: digital thermometer and infrared thermometer; two in the room or one in each room: Leslies cube, radiation sensor, ohmmeter, millivoltmeter, and window glass. INTRODUCTION In the first activity, the temperatures on the outer surfaces of the model structure will be measured using an infrared thermometer. By also using the inside and outside ambient temperatures and the known power of the heat source, the R-values for the different walls and the roof will be calculated. In the second activity, the thermal radiation from the different surfaces of a Leslies cube, which have the same temperature, will be measured and compared using a radiation sensor. BACKGROUND In the last experiment, we learned about the concepts of heat flow and thermal resistance. In this experiment we will apply this knowledge to a realistic case. We will then study thermal radiation in the second part of this experiment. Remember that the thermal power P (heat flow per unit time) through a layer of material depends on the temperature difference T across the layer and the thermal resistance R of the layer. The more the temperature difference is the more the thermal power, and the more the thermal resistance is the less the thermal power. Therefore we have

P=

T . R

(1)

When you buy layers of materials used in building structures, they usually specify the R-value Rv . The more the surface area A of the layer of material is the less the thermal resistance R . Therefore we have

R=

Rv . A

(2)

Experiment 12 12-2

When several layers of materials are stacked on top of each other, the thermal power P flows equally through each material after the steady state is reached. In this case the different surfaces will be at different temperatures and Equation 1 will still apply. If one knows the R-value Rv for each layer, the effective R-value could be found by adding the R-values. Note that the air also acts as a layer, on both inside and outside surfaces. The R-value of the layer of air, like any other material, depends on the thickness of the layerthe more the thickness, the more the R-value. PROCEDURE Activity 1: case study The heat lamps should have been left on overnight, with the model structure now being in steady state, with the temperatures being constant and the heat flow being uniform. We will calculate the R-values as follows. We will neglect the inside layer of air but only consider the outside layer. The reason is technicalwe cant measure the temperatures at the inside surfaces without opening the structure. Then we will first determine the thermal resistance Rair of the outside layer of air. Next the thermal power Pwall through each wall and the roof will be found. Finally the thermal resistance Rwall for each wall will be calculated. Note that this will include the thermal resistances of the layer of the material and inside layer of air. Read the inside ambient temperature Tambient inside from the glass thermometer inside the structure. Then measure the outside ambient temperature Tambient outside (the temperature of the laboratory room) using the digital thermometer. Record them below.

Tambient inside = _ _._ C

Tambient outside = _ _._ C

The next step is to measure the temperatures on the four lateral S D : S = 9 :1 surfaces and top surface. According to the very fundamental physical phenomenon known as blackbody radiation, the D thermal radiation from the surface of an object is almost entirely determined from its surface temperature. (The emissivity e , Figure 1. Distance-to-spot ratio D : S . D is the distance which is a material-dependent factor, is usually fixed at 0.95.) to the surface from the The infrared thermometer works by measuring the amount of the infrared thermometer and S is infrared radiation from the surface of the object within a solid the diameter of the spot on the angle given by the distant-to-spot ratio 9:1 (see Figure 1). The surface, of which the average device is calibrated in the factory to display the corresponding temperature is detected. temperature at the intensity of radiation detected. Note that the frequency spectrum of the radiation emitted from the object depends on its temperature. For most hot object, this is in the infrared range. For hotter objects, such as the filament of a light bulb, sun, etc., the radiation spectrum will also include visible light and beyond. CAUTION: Do not look into the infrared thermometer with the trigger

Experiment 12 12-3

pressed. This turns on the guide laser beam, which may permanently damage your eyes. Also do not point the device toward anyones eyes. Since the temperature on the surfaces is not uniform, you will take four measurements for each surface and calculate the average. Point the infrared thermometer from a few inches to the surface and press and hold the trigger. Record the four readings in Table 1, taken at each four quadrants of the rectangular surface (see Figure 2). After you measure the temperatures, find the average temperature Touter surface on the outer surface for each wall and record it in Table 1.

Figure 2. Take four temperature readings on each surface.

Finally we need to calculate the temperature differences between the outer surfaces and the outside as well as the inside. Calculate Touter surface ambient outside = Touter surface Tambient outside and

Tambient inside outer surface = Tambient inside Touter surface for all surfaces and enter them in Table 1.
Table 1. Outer-surface temperatures and inside and outside temperature differences.

Touter surface 1 Touter surface 2 Touter surface 3 Touter surface 4 Touter surface Touter surface ambient outside Tambient inside outer surface
C
Front (glass) Left (SHEETROCKairDUROCK) Right (SHEETROCKfiberDUROCK) Back ( DUROCK) Top (SHEETROCKfiberMasonite)

_ _._
_ _._ _ _._ _ _._ _ _._

_ _._
_ _._ _ _._ _ _._ _ _._

_ _._
_ _._ _ _._ _ _._ _ _._

_ _._
_ _._ _ _._ _ _._ _ _._

_ _._
_ _._ _ _._ _ _._ _ _._

_ _._
_ _._ _ _._ _ _._ _ _._

_ _._
_ _._ _ _._ _ _._ _ _._

ANALYSIS Since our structure is in the thermal steady state, the heat flow is uniform from the inside to the outside, meaning that the thermal power entering the walls is the same as the thermal power leaving the walls. Hence the temperatures of the walls dont change over time. Our first task is to find out how much thermal power flows through each wall and the roof (we will neglect the floor). We will assume that the thermal resistance Rair of the outside layer of air is the same for each wall and the roof. This is a fair assumption since they all have nearly the same surface area. Using Equation 1 for each wall and summing up, we have

Experiment 12 12-4
front left right back top Ptotal = Pwall + Pwall + Pwall + Pwall + Pwall

front left right Touter surfaceambient outside Touter surface ambient outside Touter surfaceambient outside + + Rair Rair Rair

(3)

back top Touter surface ambient outside Touter surfaceambient outside . + Rair Rair

Question 1.1. We know that the total thermal power Ptotal = 500. W (the total electrical power of our heat lamps). By substituting into Equation 3 this and the temperature differences from Table 1, solve for Rair .

Rair = _._ _ _ _ C / W
Question 1.2. Now find the thermal power Pwall through each wall and record it in Table 2. (See Equation 3.)
Table 2. Analysis.

Pwall
W
Front (glass) Left (SHEETROCKairDUROCK) Right (SHEETROCKfiberDUROCK) Back ( DUROCK) Top (SHEETROCKfiberMasonite)

Rwall
C / W
2

Rv wall
m C / W

Rv wall
US

_ _ _._ _ _ _._ _ _ _._ _ _ _._ _ _ _._

_._ _ _ _ _._ _ _ _ _._ _ _ _ _._ _ _ _ _._ _ _ _ _._ _ _ _ _._ _ _ _ _._ _ _ _ _._ _ _ _ _._ _ _ _

_._ _ _ _._ _ _ _._ _ _ _._ _ _ _._ _ _

Question 1.3. Applying Equation 1 to the thermal power flowing from the inside to the outer surface, we have Pwall = Tambient insideouter surface / Rwall for each wall. Using this separately for each

wall, calculate its thermal resistance Rwall and record it in Table 2. Note that the thermal resistance Rwall is the sum of the thermal resistances of the wall and inside layer of air.

Question 1.4. From Equation 2 the R-value Rv = RA , where A is the surface area. Using

Awall

0.75 m 2 , calculate the R-value Rv wall for each wall.

Question 1.5. Finally in the last column, convert the R-value Rv wall for each wall to British units used in US for commercial materials. For the conversion use

Experiment 12 12-5

1 m2 C / W = 5.6746 ft 2 F /(Btu / h) .
Question 1.6. Which wall materials result in the best thermal insulation? Does the air gap used in the left wall make a good insulator? Would you rely on DUROCK as a thermal insulator? Write down here any other comments along with what useful knowledge you gained in this experiment. Activity 2: emission and absorption of thermal radiation

In the first activity of this experiment, you used the infrared thermometer to measure the temperature of a surface. What the infrared thermometer actually measures is the intensity of the thermal radiation. Since the infrared thermometer doesnt know the actual physical properties of the surface, it assumes that it has an emissivity of e = 0.95 , which is true for most materials but not all. The total radiation emitted or absorbed by a surface is given by

P = eAT 4 ,

(4)

where = 5.6703 108 W / m2 / K 4 , e is the emissivity, A is the surface area, T is the temperature, and Equation 4 is known as the StefanBoltzmann law. For an ideal emitter, also known as a blackbody, the emissivity e = 1 . Most black materials have good emissivities, which emit or absorb the thermal radiation efficiently. White materials have smaller emissivities and shiny metal materials have the poorest emissivities. Such materials dont efficiently emit or absorb thermal radiation and can be used to insulate radiant heat, such as for thermos-bottle or window coatings. In this activity you will be measuring the thermal radiation from different surfaces of a Leslies cube. All surfaces of the Leslies cube are made of aluminum metal, therefore having good thermal conductivities. This means that all surfaces are approximately at the same temperature. The temperature will be measured by a thermally varying resistor, called a thermistor. The thermistor is embedded inside one of the surfaces of the cube and you will read/calculate the temperature from the chart attached to the cube. The thermal radiation from each surface will be measured using a radiation sensor, which measures the intensity of the infrared radiation.
Important notes: 1. It's helpful to preheat the Leslies cube at a setting of 5.0 for 20 minutes at the beginning of the laboratory period. 2. Parts 1 and 2 of this activity can be performed simultaneously. Make the measurements in Part 2 while waiting for the Leslies cube to reach thermal equilibrium at each of the settings in Part 1. 3. When using the radiation sensor, always shield it from the hot object except for the few seconds it takes to actually make the measurement. This prevents heating of the

Experiment 12 12-6

thermopile in the sensor, which will change the reference temperature and alter the reading. The shutter-lock ring on the sensor should be slid all the way back so that the sensor shutter will remain closed unless it is pressed open. Part 1: radiation rates from different surfaces

1. Connect the ohmmeter and millivoltmeter as shown in Figure 3. 2. If the Leslies cube is preheated, just set the switch to 5.0. If not, turn on the cube and set the power switch to HIGH. Keep an eye on the ohmmeter reading. When it gets down to about 40 k, reset the power switch to 5.0. 3. When the cube reaches thermal equilibriumthe ohmmeter reading will fluctuate around a relatively fixed valueuse the radiation sensor to measure the radiation emitted from each of the four surfaces of Figure 3. Leslies cube setup. the cube. Place the sensor so that the posts on its end are in contact with the cube surface. (This ensures that the distance of the measurement is the same for all surfaces.) Record your measurements in Table 3. Also measure and record the resistance of the thermistor. Use the chart on the base of the cube to determine the corresponding temperature. If the measured resistance value is near the middle between two values on the chart, take the average of the two temperatures corresponding to these values. 4. Increase the power-switch setting, first to 6.5, then to 8.0, then to HIGH. At each setting wait for the cube to reach thermal equilibrium, then repeat the measurements of Step 3 and record your results in Table 3.

Experiment 12 12-7 Table 3. Experimental data for the Leslies cube. Temperature setting Powerswitch setting 5.0 6.5 8.0 HIGH Thermistor resistance (k) Temperature ( C ) Thermal-radiation-sensor reading off the surface (mV) Black White Polished aluminum Dull aluminum

Question 2.1.1. List the surfaces of the Leslies cube in order of the amount of thermal radiation emitted. Is the order independent of temperature?

Question 2.1.2. It is a general rule that good absorbers of radiation are also good emitters. Are your measurements consistent with this rule? Explain.

Part 2: thermal radiation by common objects

Use the radiation sensor to examine the relative magnitudes of the radiation emitted from various objects around the room. (Also record the sign of the reading since the radiation sensor might show negative values.) In the space below, make a table summarizing your observations. Make measurements that will help you to answer the questions listed below.

Experiment 12 12-8

Question 2.2.1. Do different objects, at approximately the same temperature, emit different amounts of radiation?

Question 2.2.2. Can you find materials in your room that block thermal radiation? Can you find materials that don't block thermal radiation? (For example do your clothes effectively block the thermal radiation emitted from your body?)

Part 3: absorption and transmission of thermal radiation

Place the sensor approximately 5 cm from the black surface of the Leslies cube and record the reading. Place a piece of window glass between the sensor and the bulb. Does window glass effectively block the thermal radiation? Repeat with other materials.
Question 2.3.1. What do your results suggest about the phenomenon of heat loss through windows?

Question 2.3.2. What do your results suggest about the greenhouse effect?

Instruction Manual
Manual No. 012-08107

DataStudio Starter Manual


Manual No. 012-08107

Manual No. 012-08107

DataStudio Starter Manual

Introduction
What is DataStudio?
DataStudio is a data acquisition, display and analysis program. The software works with PASCO interfaces and sensors to collect and analyze data. You can use DataStudio to create and perform experiments in General Science, Biology, Chemistry, and Physics for all grade levels. Interfaces Depending on the type of computer, the following interfaces are recommended:
SCSI/serial port computers USB-enabled computers ScienceWorkshop interfaces PASPORT USB Link or Xplorer

DataStudio Requirements To use DataStudio, you need at least the following requirements: Macintosh - System 7.5 or higher, Free RAM: 8 Mb (16 Mb preferred), Serial, SCSI, or USB port, CD-ROM drive, 20 MB Free Hard disk space. Windows - Windows 95, 98 or NT 4.0, Free RAM: 8 Mb (16 Mb preferred), Serial, SCSI, or USB port, CD-ROM drive, 20 MB Free Hard disk space.

Using DataStudio
DataStudio collects and displays data during the experiment. Setting up an experiment is a simple matter of plugging sensors into the interface and configuring the software. DataStudio has many ways to view data, including a digit display, meter, graph, and an oscilloscope. There are several ways to use DataStudio:

1. Open a pre-configured experiment. 2. Open a pre-designed workbook. 3. Create an electronic workbook or configure an experiment.

DataStudio Starter Manual

Manual No. 012-08107

Equipment and Software Setup


Depending on the type of interface, instructions for setting up DataStudio and other equipment can vary. See the section that applies to your interface.

Starting DataStudio for the first time- PASPORT


If you are using a PASPort sensor, you can connect the equipment at any time. Details for proper connection can be found in the interface's manual or the sensor's Quick Start Card. Connecting a PASPORT Sensor should automatically launch the PASPORTAL window: These are pre-designed electronic workbooks. To use, simply highlight the desired lab and click Open Selected Workbook.

Click here to launch DataStudio and create an experiment on your own.

Manual No. 012-08107

DataStudio Starter Manual

If the PASPORTAL window does not launch, double-click the DataStudio icon on your desktop to launch the DataStudio software. When DataStudio opens, a "Welcome to DataStudio" navigator screen will appear with four options: Select Create Experiment from the startup screen. If DataStudio is already running, select New Activity from the File menu.

Use this option to create a new experiment.

Use this option to open an existing activity.

Create Experiment
Manually enter data into a table.

Open Activity
Enter a mathematical expression (e.g. y = x )
2

Enter Data

Graph Equation

DataStudio Starter Manual

Manual No. 012-08107

PASPORT Experiment Setup


Connect the desired sensor to a PASPORT interface (e.g. USB Link, Xplorer). DataStudio will automatically detect the presence of the sensor, and create an appropriate display.

The measurements available will be shown in the Summary panel. In some instances, clicking the Setup button can access additional measurements or units.

Set the sampling rate for the sensor.

Click the calibrate button to calibrate the sensor.

Select the unit of measurement.

The Experiment Setup window shows which sensors are connected to the computer. This window also shows the sampling rate for each sensor and available data types. Sensors requiring calibration will have a calibrate button, which activates the calibration menu.

Manual No. 012-08107

DataStudio Starter Manual

If you need to add a sensor that isn't connected to the interface, click on the Add Sensor button in the Experiment Setup window. A new window listing all sensors will appear, from which you can select the appropriate sensor.

DataStudio Starter Manual

Manual No. 012-08107

Starting DataStudio for the first time- ScienceWorkshop


Double-clicking the DataStudio icon on your desktop will launch the DataStudio software. When DataStudio opens, a Welcome to DataStudio navigator screen will appear with four options:

Use this option to create a new experiment. Create Experiment Open Activity

Use this option to open an existing activity.

Enter a mathematical Manually enter data into a table. Enter Data Graph Equation expression (e.g. y = x 2 )

ScienceWorkshop Experiment Setup


Click the Setup button to activate the Experiment Setup window. You will use this window to select sensors and set experimental conditions. If the software does not immediately recognize the interface, click the Change button and choose your interface from the list in the Please Choose Data Source window. The Experiment Setup window will then show the selected interface.

Manual No. 012-08107

DataStudio Starter Manual

Sensors Panel The Sensors panel lists all possible sensors. Scroll through the list to find the senor(s) for the experiment. To select a sensor, double-click the icon in the Sensors panel. The software will automatically choose the correct available port.

Now connect the physical sensor into the corresponding channel. When a sensor is selected, an icon will appear in the experiment setup window, with an arrow indicating the appropriate channel for each sensor. Double clicking the sensor icon in this window will open up the sensor properties window where you can set measurement(s), calibration and sampling rate. The available measurements are shown in the summary panel.

DataStudio Starter Manual

Manual No. 012-08107

Using DataStudio to configure experimentsall interfaces


DataStudio has a variety of tools to assist with configuring experiments. Using the Summary panel and associated functions helps further define the parameters of the experiment. The displays available provide a powerful method of data visualization. This section deals with creating data displays and describes the function of each.

Displaying Data
Summary Panel The Summary panel lists measurements currently available, any collected data in the experiment, along with displays.

To display data, the sensor or data must be associated with a display. Dragging a display type from the bottom area of the summary panel up to a sensor in the top area of the summary panel will create a display for the sensor or targeted data set. Displays can show multiple data types by dragging the sensor or data run from the data summary column into an open display. Some displays will be more useful than others depending on the sensors or experimental conditions.

10

Manual No. 012-08107

DataStudio Starter Manual

Creating a display for data


You can create or remove a display from the experiment at any time, even during data collection.

Available displays in DataStudio The following are the types of displays and a description of each:

Graph The graph display plots a sensor's data vs. time. To plot one data type against another, drag the data from the data summary (in the Summary panel) to the time axis (x-axis) of the graph. The new data type will replace time, producing an XY plot (i.e. Force vs. Position). Clicking and dragging a number on the axis will directly change the graph scale. Clicking and dragging the axis line itself will move the axis in the display window. Table The table display shows the numerical coordinates in paired columns. Digits The digits display shows the instantaneous value of the data as the experiment is running. Meter The meter display shows a pictorial representation of the data using a graphical meter. Histogram The histogram display plots data points that are lumped together in 'bins' as counts. The area of a bin is proportional to the frequency of the specified data range, or the number of times a specified measurement value has been observed. FFT The FFT (Fast Fourier Transform) displays the spectral decomposition of the data. Higher sampling rates will yield finer definition of the data's frequency spectrum. This display does not store data like other displays. It shows a 'time-slice' snapshot of the data. Oscilloscope The oscilloscope display plots a time-based graph, but like the FFT shows a 'time slice' snapshot. The data is not stored. This display is ideal for experiments using fast sampling rates.

11

DataStudio Starter Manual

Manual No. 012-08107

Workbook The workbook display is a powerful, self-contained authoring environment. Use this feature to create guided scientific inquiry or as a lab write-up tool. Workbooks can contain DataStudio displays, graphics, and text.

Setting Experiment Options

Options Use the Options button to set sampling options. Clicking the Options button in the Experiment Setup window will open the Sampling Options. Manual Sampling Tab This option is used with experiments that require selecting specific data points (instead of collecting continuous data). These data points can then be associated with a parameter that is not measured by a sensor. The associated parameter can be typed in manually. Clicking the Keep data values only when commandedcheck box will activate manual sampling mode. If the data that is kept will have associated manually input data, also check the "Enter a keyboard value when data is kept" checkbox. If the "Prompt for a value" checkbox is checked, when data is 'kept', DataStudio will prompt the user to manually input the associated data. The remaining options are for describing the manually input data. You can describe and name the data, prescribe units, as well as the numerical accuracy. Delayed Start Tab A delayed start condition causes DataStudio to monitor and not store any experiment data until a prescribed condition has been met. The condition can be based on time, or an experimental condition. Use the Delayed Start tab to select between time and data measurement, then set parameters for the start condition. Automatic Stop Tab An automatic stop condition causes DataStudio to end data collection when a prescribed condition has been met. The condition can be based on time or an experimental condition. Use the Automatic Stop Tab to select between time and data measurement, then set parameters for the start condition.

12

Manual No. 012-08107

DataStudio Starter Manual

Change The change button is used to switch PASCO interfaces. Use this button to select the appropriate interface:

13

DataStudio Starter Manual

Manual No. 012-08107

Taking Measurements
Collecting Data
Once the experiment is set up, click the Start button to begin collecting data.

Start button and timer When the Start button is clicked, it will change into a Stop button. Clicking the Stop button will stop data collection. The experiment timer displays the current timing condition; either how long data has been collected, or a countdown set by an initial timing condition.

Keep/Stop button If the experiment has been configured for Manual Sampling (see Setting Experiment Options) the start button will change into a Keep/Stop button. Pressing the Keep button during data collection will store a data point. Clicking the red square to the right of the Keep button will stop data collection.

Display and Analysis Tools


DataStudio provides a complement of features designed to aid with displaying and analyzing data. Displays may be created or closed at any time before, during or after data collection.

Scale to fit A graph, FFT, histogram, and meter display can auto-scale using the scale to fit tool. The entire display will automatically adjust the range so the data fills the display window. Zoom in, Zoom out, Zoom select The graph and histogram zoom tools change the view of the display window in order to shrink, expand, or focus in on a select portion of the data. To use the zoom select tool, click the tool then draw a box by clicking and dragging around the data area of interest. The graph will zoom in to the area you selected. The Scale to Fit button will return the data back to the optimal view for all data points. Smart Tool The Smart Tool activates a set of cross hairs that displays the coordinate data pair of a specific data point. As you get closer to a data point, the Smart Tool will "gravitate" towards the data point. The displayed coordinates appear in parenthesis at the upper right edge of the small box around the cross hairs. The smart tool may also be used to display the difference between two data points.

14

Manual No. 012-08107

DataStudio Starter Manual

Moving the Smart Tool To change the position of the cross hairs of the Smart Tool, hover your mouse cursor over the center of the smart tool until the pointer turns into two crossed double arrowheads and a hand. Drag the cross hairs of the Smart Tool to the desired location. To constrain the movement of the cross hairs to one axis, hover the pointer over the dashed line that is perpendicular to the axis you want to move along until the pointer turns into a hand. Drag the cross hairs to the new location.

Measuring change- Delta Tool The Delta Tool is a feature of the Smart Tool for measuring the change in the X and Y coordinates between two data points on a graph display. To use the Delta Tool: Drag the cross hairs of the Smart Tool to a data point. Hover the mouse pointer over one of the edges of the small box around the cross hairs until the pointer turns into a triangle and hand. Click and drag the triangle to the second data point A dashed box will appear with the selected data points at two of its corners. Along the sides of the dashed box you will see the numbers that are the difference between the coordinate values for the two points. To resize this box, click the corner of the box containing the arrowheads and drag it to a new data point. Clicking and dragging any dashed line will also resize the box (constraining movement to only one dimension).

Note Tool The note tool allows you to annotate a graph or histogram. You can also label individual data points.

Statistics Tool Statistics can be toggled on and off with this button. Pressing the drop down arrow next to the sigma symbol will display a list of available statistics.

Show Time Tool Toggles on and off the time component of the data pair in digits, meter, and table displays.

15

DataStudio Starter Manual

Manual No. 012-08107

Edit Data Tool Click on the Edit Data tool to edit data in a table. When the tool is activated, a copy of the data will appear in the data summary column. The original data set can never be modified in DataStudio. Using this tool activates the insert and delete row buttons. These are used to either insert a blank row or delete a highlighted row in a data table. Some experiments require manual data collection. This data can be entered into DataStudio for analysis. The easiest way to enter data is to create an empty data table. From the Experiment menu, select New Empty Data Table.

After selecting this option, the software creates a table display that is ready to be edited with new data. An indicator appears in the summary panel as well. Double clicking the measurement in the summary column will allow the data properties to be changed. Here you can set the name of the data, units and other properties such as accuracy.

Select Data / Remove Data Data can be selectively viewed by using the Data button. Data sets can be toggled on and off by clicking on the Data button, and choosing which sets are to be displayed or hidden. Data can be removed from a display by using the Remove button. Clicking the Remove button will remove the highlighted data set from the display.

Display Settings Clicking the display settings button will open a menu where display options can be changed. Clicking the down arrow next to the icon will open a menu of common display options. Double clicking in the middle of the display window will also open the display options menu.

16

Manual No. 012-08107

DataStudio Starter Manual

Calculator DataStudio incorporates a calculator feature that is capable of not only calculating mathematical expressions, but also manipulating data measurements from sensors. Similar to displays, calculations can be created or deleted at any time. The calculator may be used to graph equations, as well as perform calculations on data sets. Click the Calculate button to activate the calculator window. Enter functions in the form of y = f (x) where y = the name of the function and x = variable. Prompt DataStudio to evaluate the expression by clicking the "Accept" button. The software will highlight any undefined terms, which need to be defined before calculation can proceed. Variables can be defined as: Constant: set the variable to a numeric value. This is a local variable, and will be used only in this equation.

Experiment constant: set the variable to a numeric value that is recognized by all equations in the experiment (e.g. mass of cart = 500 g) Data Measurement: associate a data measurement with a variable. This will perform a calculation on an entire data set to convert the data into another desired quantity (e.g. calculate momentum using velocity data). Simply click, hold and drag a measurement into the calculator window and release on the variable to be defined. Model Range: define a region for the equation, as well as the number of coordinate points between the range.
Clicking on the buttons below the definition area allows terms to be selected and input automatically in the correct format. Terms are grouped as: Scientific (sin (x), cos (x), exp (x), etc.), Statistical (min (x), avg (x), etc.) Special (integral (x), derivative (x), etc.)

17

DataStudio Starter Manual

Manual No. 012-08107

Fit Tool Use the fit tool to smooth data on a graph, depending on the relationship of data types.

18

Manual No. 012-08107

DataStudio Starter Manual

Using the Workbook


The Workbook is a special container that can be used for lab writeups, or to create a lab activity. Workbooks can contain text, graphics, and data displays. DataStudio's Workbook can be used to guide a student through an activity, allowing the student to perform each step of the activity and record observations. To begin creating a workbook, doubleclick the Workbook icon in the Displays List of the Summary. A blank workbook page will open. To turn off the workbook tools, press <Ctrl> + T.

Workbook Tools

Adding a Display to the Workbook From the display summary, click and drag a display into the Workbook window. The display will appear in the window. Add a text block This tool allows you to create a text block and add text directly into the workbook. Right clicking over the text block brings up a list of options to format the text. Selecting Always Editable allows students to type in the box.

Add a Text File This tool allows you to import a text file directly into the workbook.

19

DataStudio Starter Manual

Manual No. 012-08107

Add a Picture This tool allows you to import a .bmp or .pic graphic into the workbook.

Delete Selected Item Removes a selected item from the workbook completely.

Add Page / Delete Page This tool will add a new blank page to your workbook, or delete the current page.

20

Appendix B B-1

Appendix B Error analysis OBJECTIVE To provide a brief review of the concepts involved in error analysis. To learn about error propagation in calculations. BACKGROUND Error analysis provides the underlying basis for confidence in scientific experiment. This arises from the fact that measurements cannot be exact. With any measurement there is some uncertainty introduced by limitations in equipment precision. In addition uncertainty can be introduced by an experimental method or by the experimentalist herself. It is just this task of learning how to account for sources of uncertainty in measurements that embodies error analysis. In addition to errors and uncertainties that arise from direct measurements, errors are also compounded when measurement values are used in mathematical formulas. Therefore, it is essential to know how errors interact with each other as well. PROCEDURE You will be expected to do an error analysis for most experiments. There are two basic ways to estimate the measurement errors. The first and preferred procedure is to make many measurements to obtain the mean as an expected value of the measurement and the standard deviation as a measure of the error. Alternately, if this is not practical, you may estimate your error in a single measurement. If you use the estimation method, you must justify your assumptions and choices in arriving at a value for the error. In this case, if you are uncertain, you may take the one-half the minimum division of the measuring device as the error. For example, with a meter stick with 1-millimeter divisions, this would be 0.5 mm. Notation For any experimental measurement, the result that is listed must have an estimate of error associated with it. This estimate is generally written as a range of values in which the correct answer could lie. For example a measurement of the length of a board to be 56.2 0.3 cm means that the length is somewhere between 55.9 cm and 56.5 cm. In this case the figure 0.3 cm is called the absolute experimental error. Note that the absolute error must have the same accuracy as the measured (mean) value. This means that the absolute error must have the same number of digits after the decimal point as the mean. Another value that is often useful is the relative error. Relative error is just the absolute error divided by the measured value (mean) and it expresses the experimental error as a percentage of the original stated value (in this case 0.5%). The relative error must have the same number of significant figures as the absolute error so that one can be calculated from the other and there are no redundant digits either.

Appendix B B-2

As a second example, consider an object with a measured mass of 34,700 300 grams. In this case, the zeros in 34,700 are simply placeholders and have no significance. You should display this result in scientific notation, such that there is no confusion about the significance of the zeros. (3.47 0.03) x 104 gm properly indicates that the error is 3 in the third figure of mass. Again notice that the decimal accuracy (2 positions to the right of the decimal) is identical for the result (mean) and absolute error. In addition, when making a series of measurements, the error in each measurement is unrelated to the error in the previous measurement. This type of error is referred to as random error. An additional type of error arises when the same error occurs repeatedly (i.e. someone cut the end off of your ruler). This type of error is referred to as systematic error. Random errors are statistical in nature and can be analyzed with statistical methods. Systematic error is usually discovered by making a measurement of the same physical quantity using two completely independent methods. These methods will not have the same systematic error associated with them and will therefore expose the existence of a problem. Calculation of uncertainty in a measurement A common solution to estimating the precision of a measurement is to repeat the measurement many times and assume that the average value of these measurements has a greater significance than any individual measurement. How closely the measurements agree with each other (the reproducibility of the experiment) can be used as a measurement of experimental uncertainty. If the spread of the measurements is large compared to the average value, we infer that the reproducibility and precision is poor and vice versa. Thus the width of the distribution of measurements is a measure of the uncertainty or precision of our results. The statistical method for describing a symmetric distribution is to calculate the center (the mean) and the spread (the standard deviation). For an asymmetric distribution, interpretation of the mean and standard deviation should be done with care. Specifically, the quantities are defined as:

Mean value of N measurements = x =

1 N

x
i

(1)

Standard deviation of N measurements = N 1 =

(x
i i

x )2 (2)

N 1

Utilizing this statistical method for a symmetric distribution provides a level of confidence in your measurements. The derivation of the method shows that about 68% of all measurements fall within x N 1 and about 95% of all measurements fall within x 2 N 1 . Most inexpensive scientific calculators have these statistical functions included on them and the calculation of the mean and standard deviation can be quite simple. It is advisable to do a calculation of the mean and standard deviation with a calculator and by using Equations 1 and 2 to compare the results and to make sure that you understand how the calculator works.

Appendix B B-3

Propagating errors through calculations

After making a set of measurements, it is generally desirable to use the values obtained in some mathematical formula. For example the measurement of a radius of a sphere may be used to calculate the volume of the sphere. To use the value in a formula, it is essential to know how the errors relate to each other. The basic idea is as follows. If a quantity y has to be calculated from experimentally measured quantities x1 , x2 , x3 , , which are uncorrelated with each other, one uses the formulas below to find the absolute error y in y in terms of the absolute errors x1 , x2 , x3 , in x1 , x2 , x3 , :
Addition and subtraction of uncorrelated quantities

If y = ai xi , with ai being constants (positive or negative), then


i

( y )

= ai2 ( xi ) .
2 i

(3)

Multiplication, division, and powers of uncorrelated quantities

If y = xini , with ni being constants (positive or negative), then


i

y 2 x = ni . i y xi
Question 1. Area of a square

(4)

You are given a square with two sides being almost perfectly equal. You are asked to determine its area A = a 2 by measuring its sides a and also to give a relative-error estimate A / A for its area in terms of your relative errors a / a in the measurements of its sides. If you directly use the result of Equation 4, you get A / A = 2 ( a / a ) . But if you consider the square as a rectangle with sides a b and measure both sides and use the formula A = ab , Equation 4 gives

A / A =

( a / a ) + ( b / b )
2

2 ( a / a ) = 2 ( a / a ) .
2

Should the relative error A / A for the area of the square be taken as 2 ( a / a ) or as

2 ( a / a ) ? Why or why not? If you can answer this question properly, you have understood error propagation. Hint: How does the prior knowledge that the square is almost a perfect square affect your result?

Appendix B B-4

You might also like