You are on page 1of 19

Downloaded from rsta.royalsocietypublishing.

org on April 2, 2012

Modal testing
Kenneth G. McConnell Phil. Trans. R. Soc. Lond. A 2001 359, 11-28 doi: 10.1098/rsta.2000.0711

Email alerting service

Receive free email alerts when new articles cite this article - sign up in the box at the top right-hand corner of the article or click here

To subscribe to Phil. Trans. R. Soc. Lond. A go to: http://rsta.royalsocietypublishing.org/subscriptions

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

10.1098/rsta.2000.0711

Modal testing
By K e n n e t h G. M c C o n n e l l Department of Aerospace Engineering and Engineering Mechanics, 1200 Howe Hall, Iowa State University, Ames, IA 50011, USA The basic theory of modal analysis is developed in order to evaluate the consequences of using dierent test procedures. Piezoelectric accelerometer and force transducer characteristics are reviewed in order to show their basic electrical and mechanical response characteristics. Unique force and accelerometer behaviours are reviewed as sources of unwanted signal contamination. The consequences of having 36 input/output frequency-response function relationships between any two points in a structure are reviewed in terms of nite-element program model validation, structural modication, and the combining of two substructures.
Keywords: modal; input; instrumentation; cross-axis signal; force models; future needs

1. Introduction
Experimental modal testing is used to describe the dynamic behaviour of a structure. The choice of testing procedure and instrumentation requires an understanding of what information is needed and the consequences of using particular procedures in acquiring and processing that information. It is necessary to briey review the theoretical foundation behind modal testing in order to have a road map to dene our testing requirements. This review shall be done in the simplest of terms by examining the one-dimensional vibration of a uniform simply supported beam (gure 1a), where N points are used to describe the structures motion. The motion of the pth point is shown as up (t) while the force acting on the ith point is given by Pi (t). When all of the forces are added according to Newtons second law, Rao (1995) shows that the governing dierential equation of motion is given by [m]{} + [c]{u} + [k]{u} = {P (t)}, u
inertia damping stiness excitation

(1.1)

where [m] is the symmetrical inertia matrix, [k] is the symmetrical stiness matrix, [c] is usually assumed to be proportional to [m] and [k], and the excitation vector terms Pi (t) are assumed to be independent loading functions. A special case occurs when each force vector term has the same time history f (t), so the excitation vector can be written as {P (t)} = {P }f (t). (1.2) Figure 1b shows a load distribution along the beam as described by the load vector {P }. Then equation (1.1) becomes [m]{} + [c]{u} + [k]{u} = {P }f (t). u
Phil. Trans. R. Soc. Lond. A (2001) 359, 1128

(1.3)
c 2001 The Royal Society

11

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

12
(a) up(t) xp xi Pi (t)

K. G. McConnell
(b) uN (t) Pi

Figure 1. (a) A simply supported beam showing coordinates up (t) and time-independent loading Pi (t). (b) Distributed loading vector {P } with the same time history f (t) at each point.

This second type of loading occurs when the entire structure is mounted on a vibration exciter, but is not particularly well suited for modal testing unless the results are handled carefully, since the excitation is a distributed loading on the structure. It is clear that (1.1) and (1.3) represent a dynamic force balance at each point in the structure that involves inertia, damping, stiness (or spring restoring force) and excitation forces. This equation is typical of all linear mechanical systems in that this dynamic force balance must be satised at each coordinate point in the structure for each time instant. This model uses a subset of N points to describe the behaviour of all possible points. Thus, it is limited in the amount of information that it can account for. The same limitation holds when using N experimental points and leads to spatial aliasing.

2. The linear modal model


(a) Natural frequencies, mode shapes and orthogonality The natural frequencies and mode shapes are obtained by solving the well-known eigenvalue problem (Wiley 1951) for the case when [c] = 0 and {P } = 0. This solution gives a set of undamped natural frequencies q (rad s1 ) for q = 1, 2, . . . , N that are related to the eigenvalues q with a real mode shape vector {}q for the qth natural frequency {}q . The mode shape vector {}q is the qth column of the N N modal matrix [], so [] = {}1 . . . {}q . . . {}N , (2.1) where the individual matrix element rq corresponds to the rth points modal motion corresponding to the qth natural frequency. These mode shapes have orthogonality properties with respect to the structures physical properties such that []T [m][] = [m]diag and []T [k][] = [k]diag , (2.2) where mq is the qth modal mass on the diagonal of matrix [m]diag and kq is the modal stiness on the diagonal of the stiness matrix [k]diag . Note that all o-diagonal terms are zero for [m]diag and [k]diag . It is common to assume that the damping matrix [c] is proportional to the mass matrix [m] and the stiness matrix [k]. This means that the damping matrix is also orthogonal and is given by []T [c][] = [c]diag , (2.3) where modal damping cq is the qth modal damping on the diagonal of matrix [c]diag .
Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

Modal testing

13

The qth natural frequency is related to the qth modal stiness and qth modal mass by
2 q = kq /mq .

(2.4)

This natural frequency expression is independent of how the qth mode shape is scaled since the scale factor is squared in both kq and mq and cancels in (2.4). However, the values of kq and mq depend on mode shape scaling. Hence it is common practice in both theoretical and experimental work to scale the mode shape so that the modal mass is unity. (b) Expansion into modal space The modal model is generated by making use of the orthogonality properties of a linear system as given by (2.2) and (2.3). It is assumed that the forced response vector {u} can be written in terms of a matrix transformation that uses the mode shapes and an unknown time function vector {(t)}: {u} = []{}. (2.5) This matrix transformation and the orthogonality characteristics allow us to transform from the physical space as given by either (1.1) or (1.3) into modal space, where one mode of vibration at a time is used to understand the basic elements that control the resulting motion. Now substitute (2.5) into (1.1), then premultiply the result by []T and apply the orthogonality conditions of (2.2) and (2.3). The result is [m]diag {} + [C]diag {} + [k]diag {} = []T {P }, which contains N second-order dierential equations of motion described by
N

(2.6)

mq q + cq q + kq q =
r=1

rq Pr (t)

(2.7)

for the case where each point has an independent excitation force. It is clear from (2.6) and (2.7) that the orthogonality properties allow us to decouple the multipledegree-of-freedom system into N -single-degree-of-freedom (SDOF) systems in modal space that can be worked with one at a time. (c) Single-point excitation The most common modal testing arrangement excites the structure one point at a time in order to obtain the required modal data. When the excitation force is at the ith point and all other forces are zero, (2.7) becomes mq q + cq q + kq q = iq Pi (t). (2.8) It is clear from (2.8) that the qth mode is not excited when the ith point is a node point, so iq = 0. Thus, the choice of excitation is important in understanding the resulting motions. The corresponding dimensionless modal damping ratio q is given by cq , (2.9) q = 2 kq mq where again the modal scaling cancels out.
Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

14

K. G. McConnell (d ) Solution in modal space

Equation (2.8) is a second-order linear dierential equation that is identical in form to that for an SDOF system. The steady-state solution can be easily obtained by using phasors to convert from a dierential equation in the time domain to an algebraic equation in the frequency domain. This is accomplished by assuming that Pi (t) and q (t) are phasors given by and q (t) = Bq ejt , (2.10) where is the excitation frequency, j = 1 and Bq is a complex quantity containing both magnitude and phase information. Substitution of (2.10) into (2.8) yields a standard single degree of freedom, forced vibration response of Bq = iq Pi . kq mq 2 + jcq (2.11) Pi (t) = Pi ejt

If (2.11) is substituted into (2.10) and (2.10) is substituted into (2.6), then the modal model for the displacement at location p becomes
N

up (t) =
q=1

pq iq Pi ejt = Hpi ()Pi ejt , kq mq 2 + jcq

(2.12)

where Hpi () is the receptance frequency-response function (FRF) given by


N

Hpi () =
q=1

Up () pq iq = , 2 + jc kq mq Pi () q

(2.13)

where Up () is the frequency spectrum of the response up (t) at location p and Pi () is the frequency spectrum of the input time history Pi (t). A separate paper will discuss the type of processing that is used to generate these frequency spectra for various types of time history. Hpi () is the response at location p due to a unit excitation force at the ith location which has frequency . When p and i are dierent, Hpi () is a transfer receptance. When p = i, the Hpp () is a driving-point receptance. It is clear in (2.13) that natural frequencies, modal damping and mode shape information are contained in the receptance function. Note that the units of Hpi () are length per unit force. (e) Mobility and accelerance Instead of measuring displacement, experimental modal analysis prefers to measure either velocity (primarily from laser-based sensors) or acceleration (from accelerometers). The corresponding frequency domain FRF functions are called mobility Ypi (), which relates output velocity frequency spectrum Vp () for location p due to input force frequency spectrum Pi () for location i, and accelerance Api (), which relates output acceleration frequency spectrum ap () at location p to input excitation frequency spectrum Pi () at location i. The mobility Ypi () is related to the receptance Hpi () by Ypi () = jHpi () =
Phil. Trans. R. Soc. Lond. A (2001)

Vp () , Pi ()

(2.14)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

Modal testing

15

and the accelerance Api () is related to the mobility Ypi () and the receptance Hpi () by Api () = jYpi () = 2 Hpi () = ap () . Pi () (2.15)

It is clear from (2.13)(2.15) that the pth mode shape qp is the dominant term in these FRFs near the qth natural frequency q . Hence knowledge of the natural frequency, damping and corresponding mode shape gives us insight into the importance of each mode of vibration in complicated vibratory motion, one mode of vibration at a time. (f ) The test procedure implications of Hpi () The required modal test procedure is contained in the receptance FRF Hpi (); this contains the product of two mode shape terms pq and iq and is a part of both the mobility and accelerance FRFs. The rst mode shape term pq corresponds to the response at the pth location for the qth natural frequency. Hence this mode shape term is associated with where the measurement transducer is located. The second mode shape term iq corresponds to the ith force input location. Both of these mode shape terms have the potential to describe the modal matrix [], but the choice of test procedure can signicantly inuence the measured results. These two test procedures are called the roving hammer and the roving accelerometer. For the roving hammer, let the motion transducer be attached at the pth location while the ith excitation point changes from 1 to N as the tests are run. This generates a family of FRFs with the same value of p. It is clear that all the mode shape information is contained in the iq term. The problem is that one or more of the pq = 0 for particular values of q. When this happens, there is no information about the corresponding modes in the measured data since there is no motion at that frequency at the pth location. For the roving accelerometer, let the excitation point be xed at the ith location while the motion-measuring transducer at location p is moved from 1 to N over the structure. All modal information is contained in the pq term. However, the same problem occurs when the excitation mode shape term iq = 0 for one or more natural frequencies so that the excitation is at a node point. Again, the corresponding modal information is lost. This missing modal information presents a real problem when dealing with an unknown structure, since the positions of the node points for each natural frequency are also unknown. Consequently, a set of preliminary tests is required in order to determine what natural frequencies are present and to select one or more excitation or response locations that can be used to obtain all relevant modal data. (g) Multiple-point excitation with a common time history Consider the case where multiple independent excitations occur. Then (2.15) gives the frequency domain response for each input. Since this is a linear system, these responses can be summed together to obtain the acceleration frequency spectrum as
N

ap () =
i=1

Api ()Pi ().

(2.16)

Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

16

K. G. McConnell

It is clear from (2.16) that it is dicult to extract modal information since all accelerance and force spectrum information is mixed together. However, if the conditions of (1.2) hold, then (2.16) reduces to
N

ap () =
q=1

2 pq Qq F (), kq mq 2 + jcq

(2.17)

where Qq is the qth modes excitation term that is given by


N

Qq =
i=1

iq Pi ,

(2.18)

and F () is the frequency spectrum of the common time history f (t). From (2.18) it is clear that certain load distributions and mode shapes do not correlate with one another, so Qq = 0. Hence it is possible to have an excitation condition where certain modes are not excited and therefore cannot be measured in that testing arrangement.

3. Example of two test types


These ideas are illustrated using the simply supported beam in gure 1a. It is assumed the beam has a mass of 2.0 kg, and a length and cross-section so that the fundamental natural frequency is 20.5 Hz. The simply supported beam mode shapes are given by p (x) = sin(px/l) (3.1) from the continuous model for the pth mode. The maximum value is 1; the rst mode shape is a half-sine wave, the second is a full sine wave and the third is a 1 1 -sine wave, as shown. The nine positions on the beam are used to describe its 2 motion (gure 2a) so that the discrete mode shapes {}q are the discrete amplitudes from each location for the qth mode. For nine node points, the mass entry at each diagonal position in the mass matrix [m] in (1.1) is 0.20 kg. The pth natural frequency is related to the rst natural frequency by fp = p2 f1 , (3.2) so the beams natural frequencies in hertz are 20.5, 82.0, 184.5, 328, 512.5, etc. Each mode is assumed to be lightly damped and to have the same damping ratio of 0.60%. The modal spring constant, kq , is calculated from (2.4). Then the accelerance modal FRF, Hp (), in (2.15), (2.16) and (2.17) depends only on the modal stiness, damping and inertial properties for a given frequency and is given by Hp () = 2 , kp mp 2 + jcp (3.3)

which is the response in acceleration per unit force (m s2 N1 ). The rst four modal FRFs are shown in gure 2b, where the rst peak corresponds to the rst natural frequency, and so forth. The fourth resonance occurs o the scale, since the frequency range of the proposed test covers only the frequency range 0300 Hz. However, this mode has signicant values in the upper frequency range of interest. Note that each
Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

Modal testing
(a) 1 mag. (m s2 N1) magnitude 100 10 1 0.1 0.01 0 100 200 frequency (Hz) (b)

17

1 0 2 4 6 beam position 8 10

300

Figure 2. (a) Plot of the rst three mode shapes and (b) the rst four modal FRFs for a simply supported beam.

modal accelerance FRF approaches a value of 1/mp at frequencies considerably above that modes natural frequency. In this example 1/mp = 1.0. The mode shape information from gure 2a and the modal FRFs from gure 2b are used to calculate the beams modal responses under two types of common tests. The rst test is one where a concentrated excitation force is applied at position i and the response is measured at position p; see equation (2.15). The second test is one where the beam is mounted on a vibration exciter so that an inertial distributed excitation force excites the structure while responses are measured at each position p; see equations (2.17) and (2.18). (a) Single input force: output acceleration test arrangement In this demonstration case, the input force is applied at location i and the output acceleration is measured at location p. By combining (2.13), (2.15) and (3.3), the transfer accelerance becomes
N1

Api () =
q=1

Hq ()pq iq ,

(3.4)

where the resultant FRF curve is dependent on the value of N1 . The test structure has a high value of N1 since it uses all modes. However, any model that is used to t this response is limited by the number of modes that are used to describe the phenomenon as well as the frequency range used in the measurements. To illustrate this, let p = 4 and i = 7 in (3.4); three dierent accelerance plots A4,7 () have been calculated and are shown in gure 3a. The solid curve uses all nine modes and represents a good estimate of the actual accelerance. The chain-dotted curve uses the rst four modal terms, including the fourth mode that is slightly out of the test frequency range. The dotted curve uses only the rst three modal terms. It is clear from gure 3a that any attempt to represent the actual accelerance FRF can be in considerable error unless the analysis includes a residual mode, out of the test range. This out-of-range mode usually has the greatest eect on the valleys and the least eect on the peaks, as shown here. A separate paper will address these issues in detail. The original experimental data are usually in analogue form; it is converted to digital form by an analogue-to-digital converter (ADC) and processed in a digital
Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

18
100 (a) mag. (m s2 N1) 10 1 0.1 0.01 0 50

K. G. McConnell

100

150 frequency (Hz) mag. (m s2 N1) 100 (c) 10

200

250

300

mag. (m s2 N1)

100 (b) 10 20.5

1 16

18

20 22 frequency (Hz)

24

78

80

82 84 frequency (Hz)

86

Figure 3. (a) Comparison of transfer accelerance A4,7 () calculated using 3, 4 and 9 modes. (b) The rst resonant peak showing the eect of digital representation of the FRF when the peak occurs between two frequency components. (c) The second resonant peak when one of the frequency components corresponds to the resonant peak.

frequency analyser. When these data are presented as an accelerance plot, it is done in terms of discrete, equally spaced frequencies. In this example, the frequency spacing is f = 1 Hz. Thus, the 20.5 Hz peak of the rst mode lies between two discrete frequency components of 20 and 21 Hz (gure 3b) while the 82.0 Hz peak of the second mode corresponds to one of the digital frequency components (gure 3c). It is evident that using a simple peak reading can lead to considerable measurement error in determining mode shapes and modal damping values through a lack of resolution in the frequency domain. For example, the true peak at 20.5 Hz in gure 3b is 64.1, while the two peaks at 20 and 21 Hz are 14.7 and 15.9, respectively. Thus, this peak can be easily interpreted about 4.2 times too small. This resolution problem is more serious at the lower frequencies than at high frequencies. In addition, signal noise can contribute to peak uncertainty. A separate paper will address the modal parameter extraction issues in detail. (b) Distributed excitation source: the vibration exciter test If the simply supported beam is mounted on a vibration exciter and driven with a base motion y(t), the absolute beam motion up (t) at point p is related to the exciter motion and the relative motion vp (t) between the beam and its corresponding rigid body position by up (t) = vp (t) + y (t).
Phil. Trans. R. Soc. Lond. A (2001)

(3.5)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

Modal testing
Table 1. Modal loading coecient Qq mode 1 2 3 4 uniform load 1.263 0 0.393 0 triangular load 0 0.769 0 0.344

19

The corresponding acceleration transmissibility ratio becomes


N1

Ap () = 1 +
q=1

2 Hq ()pq Qq ,

(3.6)

where Qq is the modal loading coecient. When the base acceleration motion is uniform along the length of the beam, the modal excitation force Quq becomes
9

Quq =
i=1

iq mii ,

(3.7)

where mii = 0.2 kg for each point in this case. However, if the exciter table has a pure rocking motion, the inertial excitation is a linear function with a triangular shape, so (3.7) becomes
9

Qtq =
i=1

iq mii [1 0.25(i 1)].

(3.8)

The resulting modal excitation for these two cases is given in table 1. Table 1 clearly shows that the uniform excitation loading does not excite the even modes of vibration, while the triangular loading distribution does not excite the odd modes. Note that in real life the exciter table may produce a frequency-dependent combination of linear and rocking motion that depends on the systems mass distribution and the exciters armature suspension system. Figure 4a shows the acceleration transmissibility ratio at the fourth point, A4 (), for a uniform load distribution. This transmissibility ratio is calculated from (3.6) using three, four and nine modes. The plot shows only the two odd modes of vibration since both the second and fourth modes are not present, due to the corresponding zero modal loading coecient. Consequently, there is no fourth-mode residual term in this case. Figure 4b shows the acceleration transmissibility ratio at the fourth point, A4 (), for the triangular load distribution. It shows only the eects of the two even modes since the odd modes are not excited in this case. Consequently, there is a strong residual term from the fourth mode that signicantly inuences the higher-frequency part of the curve and helps to give a nearly perfect t with the nine-term model. From (3.6) it is evident that the mode shape information is contained in the summation term that represents the relative motion. The mode shape is available if the accelerometer is moved from point to point in this case. However, it is also clear that the unity value is contaminating the mode shape information. Thus, subtracting unity from the measured FRF leaves only the summation term, and the
Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

20
mag. (m s2 N1) 100 (a) 10 1 0.1 100 (c) 10 1 0.1 0.01 0 100 200 frequency (Hz)

K. G. McConnell
100 (b) 10 1 0.1 0.01 100 (d) 10 1 300 0.1

mag. (m s2 N1)

100 200 frequency (Hz)

300

Figure 4. Acceleration transmissibility ratio A4 () for (a) uniform inertial loading and (b) triangular inertial loading. Relative acceleration transmissibility ratio A4 ()rel for (c) uniform inertial loading and (d) triangular inertial loading.

corresponding mode shape information is not contaminated by this oset. Then the FRFs, A4 (), from gure 4a, b take on the relative FRF form of A4 ()rel and are shown in gure 4c, d. Clearly, the fourth mode has signicant eects on the relative acceleration transmissibility results as well. (c) Summary This example has shown that the results depend on the test method employed. Regardless of the test method, the potential exists to completely miss a given natural frequency and mode shape. The single-input single-output method will have a missing mode if the stationary transducer has a node point for that particular mode of vibration. In the distributed excitation case, a mode is missing if the loads spatial distribution has zero correlation with that mode shape.

4. Basic instrument characteristics


An important consideration in experimental modal analysis is the quality of data that are used to extract the natural frequency, modal damping, and mode shape modal parameters. The two measured quantities used in most modal studies are acceleration for the output motion and force for the excitation input. The most common sensor uses the piezoelectric phenomenon. The main mechanical and electrical transducer characteristics are summarized here. Greater detail is available in Dally et al . (1993) and McConnell (1995). (a) The mechanical mode The simplest mechanical model for both the force transducer and an accelerometer is shown in gure 5a. The dierential equation governing this model is m + cz + kz = f (t) m, z x
Phil. Trans. R. Soc. Lond. A (2001)

(4.1)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

Modal testing
(a) F(t) y seismic mass m C c k x q R E R1 E0 (b) 1 C1

21

base 100 (c) HmHe 10 1 0.1 0.20 1.0

0.2

0.4

0.6 0.8 frequency ratio

1.0

1.2

1.4

Figure 5. (a) Basic seismic mechanical model; (b) built-in voltage follower model of a piezoelectric sensor; (c) typical electromechanical instrument FRF showing AC coupling at low frequencies and mechanical resonance at high frequencies.

where m is the seismic mass, c is the damping of the sensing crystal, k is the spring rate of the sensing crystal, z = y x is the relative motion between the seismic mass and the base, f (t) is an external load to be sensed, and x is the base motion. The corresponding mechanical FRF is given by Hm () = 1 1 = , k m 2 + jc k(1 r2 + j2r) (4.2)

where r = /n = f /fn is the dimensionless frequency ratio and is the dimensionless damping ratio. Since these instruments are designed with high natural frequencies and light damping, they are usable up to a frequency ratio of 0.2 while creating a 5% magnitude error. Hence it is evident from (4.1) that the transducers relative motion is directly proportional to external force f (t) and inertial force m, so x z 1 [f (t) m]. x k (4.3)

Equation (4.3) shows that force and pressure transducers are also sensitive to base acceleration. (b) Charge sensitivity model The piezoelectric crystal generates an electric charge that depends upon its crystal orientation and size. The charge generated is q; it is directly proportional to the
Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

22 relative motion, so q = Sz z =

K. G. McConnell Sz m Sz f (t) x = SF f (t) Sa x, k k

(4.4)

where Sz is the displacement charge sensitivity, SF is the force charge sensitivity in picocoulombs per unit of force, and Sa is the acceleration charge sensitivity in picocoulombs per unit of acceleration. (c) Basic piezoelectric electrical circuit Two basic types of piezoelectric circuits are used today. One is the built-in voltage follower and the other is the charge amplier. Figure 5b shows a simple model of the built-in voltage follower type. This model has the charge generator q connected to a capacitor C and resistor R. The output voltage E of this circuit is the input to a voltage follower amplier that has unity gain. The voltage follower output is ACcoupled through capacitor C1 to the recording instrument with an input resistance R1 . The governing dierential equation for the transducer side of the voltage follower is given by E Sq a = Sv a, E+ (4.5) = RC C where a is the time derivative of either the force or the acceleration that is being measured, E is the circuits output voltage, and Sv is the voltage sensitivity in volts per unit of acceleration or unit of force (e.g. 10.0 mV g1 or 7.50 mV N1 ) that depends on Sq and capacitance C. The FRF for the electrical circuit in gure 5b is given by He () = jRC 1 + jRC
manufacturer

jR1 C1 Sv , 1 + jR1 C1
user

(4.6)

where it is seen that the manufacturer controls the internal time constant RC and the user controls the external time constant R1 C1 by the resistance R1 that the output circuit is connected to. In contrast, the charge amplier has (a) a single internal time constant so that (4.6) has a single time constant, (b) a standardized voltage sensitivity due to the variable gain capability, and (c) a much wider dynamic range. (d ) The overall transducer mode When equations (4.2), (4.4) and (4.6) are combined, the output voltagefrequency spectrum E() is related to the input frequency spectrum a() by E() = He ()Hm ()Sv a(). (4.7) A plot of E()/Sv a() is given in gure 5c, where it is clear there is a useful working frequency range. At low frequencies it is important to match time constants so the He () terms in both the force and acceleration measurements cancel out in the FRF measurements. If the time constants are not matched, then the user must make appropriate corrections to the low-frequency end of the measured FRF using (4.6). These considerations are usually important for frequencies below 10 Hz. Thus, the usable frequency range is bounded by the RC time constant eect at the lower end and by 20% of the transducers resonant frequency at the upper end.
Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

Modal testing
//// / ///// z z x node 1 triaxial accelerometer
Figure 6. Freefree beam used for cross-axis signal contamination study: (a) coordinate locations and transducer locations; (b) impact directions Y and S.

23
S-input

side view

Y-input y node 24 end view

5. Accelerometer cross-axis sensitivity


Han (1988) and McConnell (1995) studied the eects of accelerometer cross-axis sensitivity on the coupled modes of a structure and how these coupled modes can lead to erroneous results. This cross-axis sensitivity is generally the result of manufacturing imperfections. The basic idea is that the output acceleration signal Ep for the pth direction is contaminated by the acceleration in orthogonal directions, so Sxx ax + Sxy ay + Sxz az = Ex , Syx ax + Syy ay + Syz az = Ey , (5.1) Szx ax + Szy ay + Szz az = Ez , where Spq is the acceleration voltage sensitivity for the pth direction due to acceleration in the qth direction and aq is the acceleration in the qth direction. By letting pq = (Spq /Spp ) be the cross-axis sensitivity coecient and bp = Ep /Spp be the apparent acceleration in the pth direction, then for pq bounded by approximately 5% so that second-order eects can be neglected, the actual acceleration can be estimated by ax = bx xy by xz bz , ay = yx bx + by yz bz , (5.2) az = zx bx zy by + bz . Equation (5.2) shows that triaxial acceleration measurements are required for this correction. The downside to this correction scheme is that the cross-axis sensitivity tends to change with transducer use over time due to the nature of the mechanisms that cause cross-axis sensitivity. (a) Example of cross-axis signal contamination Han (1988) ran modal experiments on a freefree steel beam that was 2337 mm long with a cross-section of 25.4 mm 28.6 mm, as shown in gure 6. The triaxial accelerometer was mounted on the left-hand end at node 1 and inputs were applied to the beam at all 24 equally spaced node points as shown. These inputs were directed either in the Y -direction or the S-direction; the Y -direction inputs should excite only
Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

24

K. G. McConnell

Table 2. Natural frequencies and modal damping values extracted from three dierent FRF sets for the same structure mode 1 2 3 4 5 6 7 Y -input fn (Hz) 27.58 75.86 148.4 244.9 y-output (%) 0.005 60 0.001 54 0.001 31 0.000 56 S-input fn (Hz) 27.59 67.55 75.91 132.3 148.5 218.1 244.9 y-output (%) 0.004 01 0.000 08 0.001 42 0.000 38 0.000 73 0.000 68 0.000 70 S-input fn (Hz) 27.58 75.91 148.5 244.9 y-output (%) 0.005 01 0.001 43 0.001 22 0.000 60

vibrations in the y-direction, and the S-direction inputs should excite vibrations in the y- and z-directions. A modal analysis based on standard analysis software was performed on each set of original data as well as the corrected S-direction data; the results are given in table 2. It is clear from this table that there are only four natural frequencies, damping ratios and mode shapes in the test frequency range for the Y -input/y-output dataset, whereas seven natural frequencies, damping ratios and mode shapes resulted from the S-input/y-output dataset. The S-input data are corrected using (5.2), where the pq were obtained from careful calibration of the triaxial accelerometer. The corrected results gave the same four natural frequencies and mode shapes, and essentially the same damping values as shown in table 1. It is clear that such signal contamination can cause signicant errors and it is clear that the data can be easily misinterpreted when dealing with an unknown structure.

6. Force transducer considerations


McConnell (1993) has shown that force transducers can be used in three basically different ways and that the transducer exhibits dierent characteristics for each type of use. In the rst way, the transducer base is attached to a rigid foundation (gure 7a). In the second way, the transducer is mounted on a hammer for impact testing (gure 7b). In the third way, the transducer is attached directly to the structure under test with a stinger between the load cell and the exciter so that the bending moments on the load cell are minimized (gure 7c). (a) Attached to rigid foundation This case is shown in gure 7a, where the base motion x in (4.1) is essentially zero and the transducer appears to behave as an SDOF system as it measures force F (t). This simple behaviour gives us false condence about the transducers behaviour in other applications. (b) Force transducer attached to a hammer This case is shown in gure 7b, where mh is the eective hammer mass and m is the eective seismic mass that includes the transducers seismic mass and any impact tips that are attached. In this case the transducers voltage sensitivity Sf is altered
Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

Modal testing
(a)
F(t )

25
(c)

(b) sensor hammer mh


m F(t ) load cell

load cell ///////////

SUT handle

exciter

stinger

Figure 7. Three general load cell environments: (a) mounted on a rigid foundation, (b) mounted on an impact hammer, and (c) attached to a structure under test (SUT).
by the eective hammer and seismic masses, so the eective voltage sensitivity Sf becomes Sf =

mh Sf . mh + m

(6.1)

This is why impact hammer manufacturers provide several voltage sensitivities for dierent combinations of hammer mass and impact tip masses. Note that the users hand mass can alter the impact hammer mass mh when using very light hammers. (c) Attaching the force transducer directly to the structure under test This case is shown in gure 7c, where it is assumed that the structure under test (SUT) has a receptance FRF of Hs (). Then the structures driving point displacement motion Yp () at location p is given by Yp () = Hs ()Fp (). jT 1 + jT
electrical

(6.2)

The seismic mass acceleration in gure 7c is [ 2 Yp ()] and (4.1) can be reduced to Ef () = Sf k [1 mHs () 2 ]F, k + jc
mechanical seismic mass

(6.3)

where Sf is the transducers voltage sensitivity, the rst bracketed term is the piezoelectric time constant eect, the second bracketed term is the mechanical eect when all inertia terms are combined with the SUTs motion Yp (), and the eective seismic mass is multiplied by the SUTSs driving-point accelerance [Hs () 2 ]. In this case seismic mass m includes all attachment hardware as well as any accelerometer mass when measuring driving-point accelerance. These seismic mass eects show that the maximum measurement error occurs in the vicinity of the SUTs natural frequencies, the very region where the best data are desired. Hu (1991) studied the eect of stingers and seismic mass on measured accelerance FRFs. He found that the force measurement errors due to [mHs () 2 ] in (6.3) can be corrected for in the frequency domain by using App () =
Phil. Trans. R. Soc. Lond. A (2001)

Hm pp () 1 mHm pp ()

(6.4)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

26
a,

K. G. McConnell
F, M q p

Figure 8. Structure with two input excitation vectors at q and two output motion vectors at p.

for the driving-point accelerance App (), where Hm pp () is the directly measured driving-point accelerance. Similarly, the transfer accelerance Apq () can be corrected for by using Apq () = [1 + mApp ()]Hm pq (), (6.5)

where Hm pq () is the measured transfer accelerance. Ewins (1986) has suggested that the base acceleration term can be compensated for in the time domain by subtracting an acceleration signal. McConnell & Hu (1993) found experimentally that such a technique works well except near resonance, when certain relative phase shift errors occur between data channels. In these cases, very large errors occurred which remain undetected. However, when (6.4) and (6.5) are used, these large errors become apparent to the user. (d ) Base strain eects Cappa & McConnell (1994) have shown how base strain can alter the measured forces due to base strain sensitivity of a compression design type of force transducer. These eects were found to depend on the test structure as well as the orientation of the transducer relative to the base strain. Similar eects are observed for compression design accelerometers when compared to shear design accelerometers. (e) Bending moment eects McConnell & Varoto (1993) have shown how bending moment sensitivity of a force transducer can contaminate the measured forces. There is no way to remove this type of contamination with a single transducer. However, Dong et al . (1998) have shown that cross-axis sensitivity between forces and moments can be removed from a single transducer that measures force and two moments, when the transducer is carefully calibrated for this cross-axis sensitivity. The ability to remove contaminated information from data due to transducer cross-axis sensitivity requires a complete set of information in order to correct the original data.

7. Future needs in experimental modal analysis


There are several potential uses for experimentally determined modal parameters and/or FRFs. Some of these uses have given less than satisfactory results. How is the modal information to be used in nite-element model verication? How should these data be used in predicting structural modications or other substructuring applications? What factors contribute to conceptual as well as measurement errors? A fundamental contributor to these problems is our concept of what constitutes a complete input/output model for a structure.
Phil. Trans. R. Soc. Lond. A (2001)

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

Modal testing

27

Consider the situation shown in gure 8, where force F and moment M are the inputs at location q, and linear acceleration a and angular acceleration are outputs at location p. Since each input and each output is a vector with three orthogonal components, there are in total six inputs and six outputs, and at each frequency this leads to 36 complex input/output relationships between these two points, complex relationships with real and imaginary parts. These relationships can be written as follows: a
p

A B C D

pq

F M

(7.1)

In equation (7.1), A is a 3 3 submatrix that is the linear accelerance between acceleration a and force F . B is a 3 3 submatrix that is the linear accelerance between acceleration a and moment M . C is a 3 3 submatrix that is the angular accelerance between angular acceleration and force F . D is a 3 3 submatrix that is the angular accelerance between angular acceleration and moment M . Reciprocity holds between B and C, so B = C T or C = B T . Current practice is usually limited to measuring the three linear acceleration components of a while using a triaxial accelerometer and typically measuring a single component of the input force F . This test procedure gives the rst column of submatrix A and ignores the remaining 33 potential input/output relationships. The eects of unmeasured inputs can be seen by expanding the pth row of (7.1) to give ap = Ap1 F1 + Ap2 F2 + Ap3 F3 + Bp1 M1 + Bp2 M2 + Bp3 M3 A F1 , (7.2) =
p1

can contain signicant contamination from unmeasured where it is clear that forces and moments. A second measurement problem is that boundary conditions in the test can be subtly dierent from those used in the nite-element model. Consequently, a comparison of experimental modal results and nite-element modal results can have signicant discrepancies. Recently Varoto (1996) developed a theoretical model called READI (rules for the exchange and analysis of dynamic information). This model uses substructuring concepts to describe the processes and information that are required to perform adequate laboratory tests based on eld vibration measurements when the test item is attached to its host vehicle at multiple points. In attempting to apply this theoretical model in the automotive industry, where four interface connections occurred, the predicted combined behaviour from measurements on the two substructures was poor at certain frequencies. It became abundantly clear that this discrepancy involved most of the terms in (7.1), since a signicant shear force as well as two orthogonal moment terms were involved at each interface point. In addition, the angular motions and stiness played a signicant role. Consequently, the use of data based on one input force and three output linear accelerations was inadequate. In reality, there were three output linear accelerations and at least two output angular accelerations along with three input force components and two input moment components. Hence this physical situation required more like 25 input/output relationships for each interface point instead of only three. This experience illustrates several valuable lessons. First, experimental methodologies and instruments are needed so that the 36 input/output FRF relationships can be eciently measured when required by the problem at hand. Second, methods and criteria are needed to judge the importance of a given FRF in a given application.
Phil. Trans. R. Soc. Lond. A (2001)

A p1

Downloaded from rsta.royalsocietypublishing.org on April 2, 2012

28

K. G. McConnell

Third, there is a need to grasp the meaning and implications of linear and angular mode shapes in describing a structures dynamics. Fourth, there is a need to learn the consequences of these linear and angular mode shapes when attempting to do structural modications. This is particularly true when moments and shear forces are important in the combined structure at the interface points but may not appear to be important while testing the substructures before assembly.

References
Cappa, P. & McConnell, K. G. 1994 Base strain eects on force transducer measurements. In Proc. Spring Society for Experimental Mechanics, Baltimore, MD, pp. 520530. Dally, J. W., Riley, W. F. & McConnell, K. G. 1993 Instrumentation for engineering measurements, 2nd edn. Wiley. Dong, J., McConnell, K. G., Alfonzo, M. & Golovanova, L. 1998 Error reduction of measured impact forces and their lines of action via cross-axis sensitivity studies. In Proc. 16th Int. Modal Analysis Conf., Santa Barbara, CA, pp. 10991105. Ewins, D. J. 1986 Modal testing: theory and practice. Letchworth, UK: Research Studies Press. Han, S. 1988 Eects of transducer cross-axis sensitivity on modal analysis. PhD thesis, Iowa State University, Ames, IA. Hu, X. 1991 Eects of stinger axial dynamics and mass compensation methods on experimental modal analysis. PhD thesis, Iowa State University, Ames, IA. McConnell, K. G. 1993 The interaction of force transducers with their test environment. Modal Analysis 8, 137150. McConnell, K. G. 1995 Vibration testing: theory and practice. Wiley. McConnell, K. G. & Hu, X. 1993 Why do large FRF errors result from small relative phase shifts when using force transducer mass compensation methods? In Proc. 11th Int. Modal Analysis Conf., Kissimee, FL, pp. 845859. McConnell, K. G. & Varoto, P. S. 1993 A model for force transducer bending moment sensitivity and response during calibration. In Proc. 11th Int. Modal Analysis Conf., Kissimmee, FL, pp. 516521. Rao, S. S. 1995 Mechanical vibrations, 3rd edn. Reading, MA: Addison-Wesley. Varoto, P. S. 1996 Rules for the exchange and analysis of dynamic information. PhD thesis, Iowa State University, Ames, IA. Wiley, C. R. 1951 Advanced engineering mathematics. McGraw-Hill.

Phil. Trans. R. Soc. Lond. A (2001)

You might also like