Sunteți pe pagina 1din 29

Vibratory Motion and Single Degree of Freedom Systems

Table of Contents SDOF Manual.


Complementary Mathcad files are listed (please find them on the course website).
Notes and files adapted from Dr. G.J. Rix and Dr. J.R. Valdes.

Item (in handout) Associated Mathcad Files

Free Vibrations Undamped Free Vibrations.mcd


Damped Free Vibrations.mcd

Forced Vibrations Forced Vibrations.mcd

Transfer Functions Transfer Functions.mcd


Ground Displacement.mcd
Ground Acceleration.mcd

Fourier Analysis (not included in handout) Fourier Analysis.mcd

Response to Arbitrary Excitations SDOF Response to Non-Harmonic Load.mcd

Multiple Degree of Freedom Systems MDOF Systems.mcd

Complex Notation
Free Vibrations

We begin by examining the response of the single-degree-of-freedom (SDOF) system with


no external forces. The mass is set into motion by an initial displacement from its at rest
position and/or an initial velocity. We will consider two systems, one with no means of
dissipating energy and another with a viscous damping in the form of a dashpot.

Undamped Free Vibrations

Consider the single-degree-of-freedom (SDOF) system


shown at the right that has only a spring supporting the
mass. If we examine a free-body diagram of the mass we m
see that the forces acting on it include gravity (the
weight) and the resistance provided by the spring. u(t)
We use (1) Hooke’s law (F = ku) and Newton’s second k
law (F = ma) to write the following:

W − k( u + δ static ) = mu
&& (1)

where δ static is the displacement of the mass when the


system is at rest. The additional component of the W
displacement, u, is measured relative to this "at rest"
position. The static displacement is related to the weight
by:
m
W = kδ static (2)

Thus, Eq. 1 simplifies to: k ( u + δ static )


&& + ku = 0
mu (3)

which is the equation of motion for the undamped SDOF system. The objective is to solve
the equation of motion to determine the displacement of the mass as a function of time,
u(t), subject to the initial conditions of the system.

One approach to solving this partial differential equation is to assume a solution of the
form:

u( t ) = Ae rt (4)
Take the second derivative with respect to time of Eq. 4 yielding:

u t ) = r 2 Ae rt
&&( (5)

and substitute Eqs. 4 and 5 into Equation 3 giving:

mr 2 Ae rt + kAe rt = 0 (6a)

which simplifies to:

mr 2 + k = 0 (6b)

Solving for r produces two possible roots:

k
r = ±i (7)
m

where i = −1 . Thus the solution is given by:

k k
i t −i t
u( t ) = Ae m + Be m (8)

We define the radical term in the exponent to be the circular natural frequency of
vibration of the system:

k
ωn = in rad/sec (9)
m

Note that the circular natural frequency is related to the natural frequency by:

ωn
fn = (10)

and the natural period by:

1
Tn = (11)
fn

To evaluate the coefficients A and B in Eq. 8, we need two initial conditions. Usually,
these are the initial displacement and velocity of the mass:

u( t = 0) = u 0 (12a)
u& ( t = 0) = u& 0 (12b)

After substituting these expressions into Eq. 8 and its first derivative, we obtain:

u iu&  u iu& 
u( t ) =  0 + 0  e − iω n t +  0 − 0  e iω n t (13)
 2 2ω n   2 2ω n 

Figure 1 shows the displacement time history of this SDOF with the initial conditions of
u 0 = 1.0 and u& 0 = 0.0 . Note that the amplitude of the displacement does not diminish with
time because there is no means of attenuating energy within the system.

Figure 1 Displacement Time History for Free Vibrations of an Undamped SDOF System
Damped Free Vibrations

Consider the single-degree-of-freedom (SDOF)


system shown at the right that has both a spring
and dashpot. If we examine a free-body diagram of
m
the mass we see that an additional force is u(t)
provided by the dashpot. The force is proportional
to the velocity of the mass. k c

Fdamping = cu& (14)

where c is the viscous dashpot coefficient.


Summing forces in the vertical direction yields: W
W − k (u + δ static ) − cu& = m&u& (15)
m
After simplifying using Eq. 2, the equation of
motion of the system is: k( u + δ static ) cu&
&& + cu& + ku = 0
mu (16a)

or

c k
&&u + u& + u = 0 (16b)
m m

To solve this differential equation, assume a solution of the form:

u( t ) = Ae rt (17)

Take the first and second derivative with respect to time of Eq. 17 yielding:

u& ( t ) = rAe rt (18a)

u t ) = r 2 Ae rt
&&( (18b)

and substitute Eqs. 18a and b into Eq. 16b giving:

c k
r 2 Ae rt + rAe rt + Ae rt = 0 (19a)
m m

which simplifies to:


c k
r2 + r+ =0 (19b)
m m

Solving for r produces two possible roots:

−c
2
 c  k
r= ±   − (20)
2m  2m  m

Thus the solution is given by:

u( t ) = Ae r1t + Be r2 t (21)

We define a critical value of c such that the term inside the radical equals 0:

c crit = 2 km (22)

and a fraction of critical damping:

c c
β= = (23)
c crit 2 mω n

Rearranging Eq. 23 yields:

c
= ω nβ (24)
2m

which can be substituted into Eq. 20 to yield:

r = − ω n β ± iω n 1 − β 2 (25)

After substituting, Eq. 21 becomes:

[
u( t ) = e −ω nβt Ae −iω n 1− β 2 t
+ Be iω n 1− β 2 t
] (26)

To evaluate the coefficients A and B in Eq. 26, we need two initial conditions. Usually,
these are the initial displacement and velocity of the mass:

u( t = 0) = u 0 (27a)

u& ( t = 0) = u& 0 (27b)

After substituting these expressions into Eq. 26 and its first derivative, we obtain:
 u i( u& 0 − ω n βu 0 )  −iω n u i(u& 0 − ω n βu 0 )  iω n 
u( t ) = e −ω n βt  + e 1− β 2 t
+  − e 
1− β 2 t
0 0
2  2 
 2 2ω n 1 − β   2 2ω n 1 − β  
 
(28)

Consider the solution for three different values of β:

1. β = 1 (critically damped)

For β = 1 Eq. 25 reduces to:

r = −ω n (29)

and the partial differential equation has repeated roots. As a result, the solution takes the
form:

[ ]
u( t ) = u 0 + ( u& 0 + ω n u 0 )t e − ω n t (30)

Figure 2 shows the response of a critically damped SDOF for three initial conditions. The
initial displacement is equal to 1.0, but the initial velocity of the mass is 1.0, 0.0, and –1.0
for the three different cases. Notice that the motion quickly diminishes to zero because of
the large damping in the system.

Figure 2 Displacement Time History of a Critically Damped SDOF System


2. β > 1 (overdamped)

The overdamped case is similar to the critically damped case. Eq. 28 can be used directly
since the roots are not repeated. Figure 3 below shows the response of an overdamped
SDOF for the same three initial conditions in Fig. 2. As in the case of the critically damped
SDOF, the displacement quickly diminishes to zero because of the large damping in the
system.

Figure 3 Displacement Time History of an Overdamped SDOF System

3. β < 1 (underdamped)

The case of most interest to us in soil dynamics problems is that in which the fraction of
critical damping is less than 1.0. Equation 28 may be used again to calculate the
displacement as a function of time. An example of a typical displacement time history for
an underdamped SDOF is shown in Fig. 4.

Notice that the displacement time history continues for many cycles of motion (i.e.
oscillatory). The motion can be considered to be a harmonic function modulated by a
decreasing exponential function. The origin of these two functions can be seen in Eq. 28.
Figure 4 Displacement Time History of an Underdamped SDOF System

It is common to define the damped circular natural frequency as:

ωd = ωn 1− β (31)

along the corresponding damped natural frequency and damped natural period, fd and Td,
respectively. For small values of β, ωd ≈ ωn.

Another widely used measure of the damping in a viscous system is the logarithmic
decrement:

1  u( t )  2πβ
δ= ln = ≅ 2 πβ for small β (32)
n  u( t + nTd )  1 − β2
Parallel and Series Springs

It will sometimes be necessary to consider springs acting in parallel or series. This is easily
handled by determining the effective stiffness of a single, equivalent spring as shown
below.

m
u(t)
k1
1 1 1
= +
k eff k 1 k 2
k2

m
u(t)
k eff = k 1 + k 2
k1 k2
Forced Vibrations

Consider a single-degree-of-freedom (SDOF) system that


is excited by an external harmonic force as shown at the Pe iΩt
right. The equation of motion now becomes:

&& + cu& + ku = Pe iΩt


mu (33)
m
u( t )
The solution of this non-homogeneous differential
equation is composed of the a general solution from the
k c
corresponding homogeneous equation and a particular
solution:

u( t ) = u general ( t ) + u particular ( t ) (34a)

Equation 26 is the general solution and may be substituted to yield:

[
u( t ) = e −ω nβt Ae −iω n 1− β 2 t
+ Be iω n 1− β 2 t
]+ u particular ( t) (34b)

Assume that the particular solution has the same form as the external force:

u particular ( t ) = Ce iΩt (35)

where C is a constant. As before we need to take the first and second derivatives with
respect to time to substitute into the equation of motion.

u& particular ( t ) = iΩCe iΩt (36)

&&u particular ( t ) = −Ω 2 Ce iΩt (37)

After substituting and factoring common terms we obtain:

(− Ω 2 m + iΩc + k )Ce iΩt = Pe iΩt (38)

We can solve for C and substitute into Eq. 35:

P
u particular ( t ) = e iΩt (39)
k − Ω m + iΩ c
2

which can also be expressed as:


P 1
u particular ( t ) = e iΩt (40)
k  Ω2 Ω 
1 − + 2iβ 
 2
ωn ωn 
 

Summing the general and particular solutions yields:

 1 2t 1 2t  P
u ( t ) = e −ωn βt Ae − ωn −β + Be ωn −β  +
1
i i
e iΩt (41)
  k  Ω 2 Ω 

1− + 2iβ
 2
ωn ωn 
 

The transient component will include the transient response of the system due to the initial
displacement and velocity of the mass and the transient response of the system due to the
sudden application of the external force. The coefficients A and B may be determined by
setting Eq. 41 and its first time derivative equal to u 0 and u& 0 , respectively. The resulting
expressions for A and B are:

 iω β + ω 1 − β 2 u  − iω β − ω 1 − β 2 + Ω 
 n n  0 iu& 0  n n 
A=  + + P 1
2
2ω n 1 − β 2 2ω n 1 − β 2 2ω n 1 − β 2 k Ω Ω
1− + 2iβ
ωn2 ωn

(42a)

 − iω β + ω 1 − β 2 u  iω β − ω 1 − β 2 − Ω 
 n n  0 − iu& 0  n n 
B=   + + P 1
2
2ω n 1 − β 2 2ω n 1 − β 2 2ω n 1 − β 2 k Ω Ω
1− + 2iβ
ωn2 ωn

(42b)

Figure 5 shows the response of a single degree of freedom system for which the external
force has a circular frequency, Ω, equal to 1 rad/sec and the initial displacement and initial
velocity of the mass are both equal to zero.

Note that during the first 20 to 25 seconds the response of the system includes both a
transient component (the general solution) and a steady-state component (the particular
solution). After about 25 seconds the transient response has diminished to near zero and
the overall response is dominated by the steady-state response. Thus for many forced
vibration problems of interest, we can neglect the transient component of the solution
since it is not significant after the first few cycles of motion.
Figure 5 Response of SDOF System to Forced Vibrations

One way to characterize the response of the SDOF system is to relate the amplitude and
phase of the displacement to the external force. Once steady-state conditions are reached,
the ratio of the amplitudes is about 1.5 (displacement) to 1.0 (force). The displacement is
about 180 degrees out of phase with respect to the displacement. Later, we will see that
examining the amplitude and phase as a function of frequency is a convenient way to
characterize a single-degree-of-freedom system.
Rotating Mass Excitation

For machinery vibrations problems involved rotating machinery, the dynamic forces are
often the result of out-of-balance or eccentric masses as shown in the figure below.

me

The dynamic force resulting from the eccentric mass is given by:

P( t ) = m e eΩ 2 exp(iΩt ) (43)

Notice that the magnitude of the force is frequency dependent unlike the constant force
excitation defined in Eq. 33.
Transfer (Frequency Response) Functions

To characterize the response of a SDOF system to forced vibrations it is useful to define a


transfer function or frequency response function between the input and output of the
system.

Output
H (Ω ) = (44)
Input

By normalizing the output of the system with respect to the input, we emphasize the
characteristics and response of the system over the characteristics of the output or input.

Let's define a transfer function between the steady-state displacement output and force
input of a SDOF system undergoing forced vibrations:

Displacement u ss ( t )
H (Ω ) = = (45a)
Force Pe iΩt

Pe iΩt
H (Ω ) = k − Ω miΩt+ iΩc
2
(45b)
Pe

1
H (Ω ) = (45c)
k − Ω m + iΩ c
2

1 1
H ( Ω) = (45d)
k Ω 2

1 − 2 + 2i β
ωn ωn

Notice that the transfer function is a complex-valued quantity meaning the response of the
SDOF system can be characterized by a magnitude and phase. Figure 6 shows the
magnitude and phase plots for a SDOF system expressed as a function of the normalized
frequency, Ω ωn .
Figure 6 Magnitude and Phase of Transfer Function

Equations 45c and 45d and Figure 6 can be used to provide insight into the parameters
that control the response of a SDOF in different frequency ranges. Note in Equations 45c
and d that when Ω→0, the transfer function reduces to:

1
H (Ω = 0) = (46)
k

Thus, the stiffness of the system controls the response at low frequency. As Ω→ωn, the
transfer function reduces to:

1 1
H (Ω = ω n ) = = (47)
iΩc 2ikβ

and the response of the system is controlled to a large extent by the damping in the
system. Finally, as Ω becomes large, the transfer function becomes:

1
H (Ω → ∞ ) ∝ (48)
− Ω2 m
and the response of the system is largely controlled by the mass (the inertia) of the system.

Other Forms of the Transfer Function

The transfer function defined above was expressed in terms of the displacement. Other
response quantities such as the velocity and acceleration of the mass can also be used to
define a transfer function for various applications. The names associated with each of
these transfer or frequency response functions are given in Table 1.

Table 1 Transfer Functions Used in Vibration Analysis (after Inman, 1994)

Response Parameter Transfer Function Inverse Transfer Function


Displacement Receptance Dynamic Stiffness
Velocity Mobility Impedance
Acceleration Inertance Apparent Mass

Ground Displacement and Acceleration

Consider the situation in which the system


vibrates because of motion introduced at the
base of the system, not by a force applied to
m u(t)
the mass. The system is shown at the right.

The forces exerted by the spring and the


dashpot on the mass are functions of the k c
relative displacement and velocity between
the mass and the base of the system. The
absolute acceleration of the mass is still used
(F=ma).

Thus, the equation of motion of the system z( t )


becomes:

&& + c( u& − z&) + k ( u − z) = 0


mu (49)

The right hand side is set equal to zero because there are no external forces applied to the
mass. Rearranging yields:
&& + cu& + ku = cz& + kz
mu (50)

We can assume that the ground motion is given by a harmonic function:

z( t ) = Ae iΩt (51)

and that the response of the mass is given by:

u( t ) = Be iΩt (52)

After differentiating, substituting, and solving, we obtain the solution for the steady-state
displacement of the mass:

k + iΩ c
u( t ) = Ae iΩt (53)
k − Ω 2 m + iΩ c

We can also define a transfer function between the displacement of the mass and the input
ground displacement:

u( t )
H (Ω ) = (54a)
z( t )

k + iΩ c
Ae iΩt
H (Ω) = k − Ω m +iΩiΩ
2
c (54b)
Ae t

k + iΩ c
H (Ω ) = (54c)
k − Ω 2 m + iΩ c


1 + 2 iβ
ωn
H (Ω ) = (54d)
Ω2 Ω
1 − 2 + 2 iβ
ωn ωn

As before we can express the complex-valued transfer function in terms of the magnitude
and phase:
Figure 7 Magnitude and Phase of Transfer Function for Ground Displacement
In many earthquake engineering problems,
the ground motion input at the base of the
system is specified in terms of a ground
m u(t)
acceleration.

z t ) = Ce iΩt
&&( (55)
k c
Furthermore, the response parameter of
interest is the relative displacement between
the base and the mass. The latter is important
because it is the relative displacements that
are proportional to the forces induced in the
structure. &&z( t )
y( t ) = u( t ) − z( t ) (56)

Recall that the equation of motion is:

&& + c( u& − z&) + k ( u − z) = 0


mu (49 again)

After substituting Eq. 56 we obtain:

m(&&y + &&)
z + cy& + ky = 0 (55a)

or

&& + cy& + ky = − mz
my && (57b)

After solving for y(t) we obtain:

−m
y( t ) = Ce iΩt (58)
k − Ω 2 m + iΩ c

Finally, we can define the transfer function between the ground acceleration and the
relative displacement of the mass and base:

y( t )
H (Ω ) = (59a)
&&(
z t)

−m
H ( Ω) = (59b)
k − Ω m + iΩ c
1

ω 2n
H (Ω ) = (59c)
Ω2 Ω
1 − 2 + 2iβ
ωn ωn

Figure 8 Magnitude and Phase of Transfer Function for Ground Acceleration


Response to Arbitrary Excitations

Thus far we have assumed that the inputs and outputs of single-degree-of-freedom
(SDOF) systems have been harmonic functions of the form:

u( t ) = Ae iΩt (60a)

p( t ) = Be iΩt (60b)

This form is convenient for deriving transfer functions for SDOF systems, but is of
limited practical value by itself because many actual inputs and outputs are more
complex. We must combine the solutions we have developed for harmonic excitation
with another tool, the Fourier Transform, to allow us to determine the response of SDOF
systems to an arbitrary excitation. It is important to note that the linear (or equivalent
linear) nature of most soil dynamics problems enables us to use this approach.

Fourier Transforms

The Fourier Transform is a way to decompose arbitrary displacement and/or force time
histories into their harmonic components. Once decomposed, transfer functions can be
applied to the individual harmonic components. Finally, an inverse Fourier Transform can
be used to “reassemble” the individual harmonic responses to obtain the response time
history.

Let's assume we have a displacement or force time history which we denote as x(t). We
can express x(t) as a Fourier series:


x( t ) = a 0 + ∑ (a n cos nΩ 0 t + b n sin nΩ 0 t ) (61a)
n =1

where

T
1
a 0 = ∫ x( t ) dt (61b)
T0

x( t ) cos(nΩ 0 t ) dt
2
T ∫0
an = (61c)
T

b n = ∫ x( t ) sin(nΩ 0 t ) dt
2
(61d)
T0

or using complex exponentials:


x( t ) = c n ∑e
n =−∞
inΩ 0 t
(62)

In a Fourier series, each term is separated by ∆f = f0. As ∆f decreases to 0, we reduce the


Fourier series to an integral:


x ( t ) = ∫ X(Ω) e iΩt dΩ (63a)
−∞

with continuous "Fourier coefficients" given by:

X( Ω ) = ∫ x( t ) e − iΩt
dt (63b)
−∞

Equation 63b is the continuous Fourier transform used to convert a time history x(t) to the
frequency domain in which the signal is portrayed as a function of frequency. Note that
X(Ω) is a complex-valued function. Similarly, Eq. 63a is the inverse Fourier transform
enabling us to go from the frequency domain to the time domain.

We typically do not record continuous or analog signals with modern digital


instrumentation. Instead we measure a continuous signal that is digitized and stored at a
sampling interval ∆t. Thus we must use a discrete Fourier transform (DFT):

N −1
X( k∆f ) = ∑ x( n∆t ) e − i 2 πk∆f∆t (64a)
n=0

and the inverse discrete Fourier transform:

1 N −1
x( n∆t ) = ∑ X( k∆f ) e i 2 πk∆f∆t (64b)
N n=0

where N is the total number of points in the time history and k and n are and frequency
and time indices, respectively.

The fast Fourier transform (FFT or IFFT) is simply a numerically efficient version of the
DFT. The algorithm used to calculate the FFT is most efficient when the number of
points being transformed is equal to an integer power of 2 (i.e., 2n where n is an integer).
Thus it is common to increase the increase the number of points to the next larger integer
power of 2 by "padding" the end of the record with a sufficient number of zeros.

There are several features of the Fourier Transform that are helpful to know when using it
as a tool to solve problems. Suppose that a force or displacement time history contains N
points (where N = 2n) with sampling interval ∆t and that the total length of the time
history is T. When the FFT algorithm is applied to that time history to transform it to the
frequency domain, the result will be N values in the frequency domain separated by:

1
∆f = (65)
T

The Nyquist frequency is the maximum frequency that can be accurately resolved by a
sampling interval ∆t and is equal to:

1
f Nyquist = (66)
2∆t

Finally, for real-valued time histories (the most common case) only the first N/2 + 1
points in the frequency domain are unique. The remaining points are complex conjugates
of values in the first half of the record. One must be extremely careful in selecting the
correct frequencies and mathematical operations for values in the second half of the
frequency domain record to obtain correct results from the inverse FFT algorithm.

Frequency Domain Solutions

Using Fourier Transforms we can decompose a complex time history into its harmonic
components. This enables us to apply the transfer functions we developed for harmonic
motion. We will use a 3-step approach:

Use the FFT to calculate the amplitude and phase spectra corresponding to a given input
time history. The input time history may be a force (active loading) or a ground motion
(passive loading).

p(t)→FFT→P(Ω)

Calculate the response of the SDOF system in the frequency domain using the transfer
(frequency response) function:

U(Ω) = H(Ω) P(Ω)


Use the inverse FFT to obtain the response of the SDOF system in the time domain.

U(Ω)→IFFT→u(t)

Time Domain Solutions (Duhamel’s Integral)

It is also possible to calculate the response of a SDOF to an arbitrary forcing function


directly in the time domain using Duhamel’s integral. Consider the following equation of
motion for a SDOF system subjected to an arbitrary force p(t):

m!u! + cu! + ku = p( t ) (67)

The solution to Eq. 67 is given by Duhamel’s integral:

e − ωn βt t ωnβτ
u (t ) = ∫e p(τ)sin (ω d (t − τ )) dτ (68)
mω d 0

where τ is a dummy variable. Interested readers are referred to Weaver et al. (1990) or
other texts on engineering vibrations for the derivation of Duhamel’s integral. Equation
68 may also be viewed as a convolution integral in which the forcing function p(t) is
convolved with the impulse response of a SDOF system.

Unfortunately, the application of Eq. 68 is limited to situations in which the parameters of


the SDOF system (k, m, and c) are independent of frequency. In many machine-
foundation vibration problems, the equivalent values of k and c are functions of the
frequency of the excitation.
Multiple Degree of Freedom Systems

We now want to extend our analysis of the response


of lumped parameter systems to include systems in m1
p1(t) u1(t)
which there are two or more degrees of freedom.
One such system is shown at the right and has two
masses, springs, and dashpots. In geotechnical
k1 c1
engineering, mutiple degree of freedom systems are
common in foundation vibration problems involving
coupled motion (e.g. rocking and horizontal p2(t) m2 u2(t)
translation), in site response problems for layered
soil deposits, and pile dynamics problems.
k2 c2
In the 2-degree-of-freedom (2DOF) system shown
here, the equation of motion for the first mass is:

m1!u!1 + c1 (u! 1 − u! 2 ) + k 1 (u 1 − u 2 ) = p1 (t ) (69a)

Note that the forces exerted by the dashpot and the spring depend on the relative velocity
and displacement between the two masses, respectively. For the second mass, the
equation of motion is:

m 2 !u! 2 − c1 (u! 1 − u! 2 ) + c 2 u! 2 − k 1 (u 1 − u 2 ) + k 2 u 2 = p 2 (t ) (69b)

We can combine the two equations of motion into a system of equations and express
them in matrix form as follows:

 m1 0   !u!1   c1 − c1   u! 1   k 1 − k 1   u 1   p1 (t )
0 + + =
m 2  !u! 2  − c1 c1 + c 2  u! 2  − k 1 k 1 + k 2  u 2  p 2 (t )
(70a)

or

!! + Cu! + Ku = p
Mu (70b)

Direct Solution of the Equations of Motion

One approach to determining the motion of the two masses is to formulate the solution
directly. Assume that the displacements of the masses and the external forces are of the
form:
u = Ue iΩt (71a)

and

p(t ) = Pe iΩt (71b)

After substituting Eqs. 71a and 71b into Eq. 70b and rearranging, we obtain:

[− Ω M + iΩC + K ]U = P
2
(72)

Solving for the vector of displacements yields:

u 
[
U =  1  = − Ω 2 M + iΩC + K ]
−1
P (73)
u 2 

We can rewrite Eq. 73 as follows:

U = BP (74a)

where

1 K 22 − Ω 2 M 22 + iΩC 22 − K 12 − iΩC12 
B(Ω ) =   (74b)
D − K 21 − iΩC 21 K 11 − Ω M 11 + iΩC11 
2

and

( )( )
D = K 11 − Ω 2 M11 + iΩC11 K 22 − Ω 2 M 22 + iΩC 22 − (K 12 + iΩC12 )2 (74c)

Equations 74b and 74c are the complex-valued frequency response functions for the
2DOF system.

Although direct solution of the equations of motion is straightforward, it has two


limitations. First, as the number of degrees of freedom increases, the calculation of the
matrix inverse in Eq. 73 for each frequency becomes more computationally expensive.
Second, the direct solution does not yield any physical insight into the characteristics of
the MDOF such as the natural frequencies, mode shapes, and modal damping for each of
the m degrees of freedom. These characteristics are more apparent in the modal solution
of the equations of motion.
Modal Analysis of the Equations of Motion

Another solution to Eq. 70b is given by the normal mode solution or modal analysis. To
obtain the normal mode solution, it is convenient to assume proportional damping of the
form:

C = β K + γM (75)

The normal mode solution is calculated by solving the generalized eigenvalue problem:

(K − ω M ) Φ = 0
2
(76)

where ω 2 is a diagonal matrix containing the generalized eigenvalues and Φ contains the
eigenvectors. Each rth eigenvalue is the square of the circular natural frequency of the rth
mode of vibration of the N degree of freedom system. The corresponding eigenvector
gives the mode shape of the rth mode of vibration.

The generalized stiffness, mass, and damping matrices are given by:

m = Φ T MΦ (77a)

k = Φ T KΦ (77b)

c = Φ T CΦ (77c)

The modal damping ratio for each mode is given by:

βω r γ
ζr = + r = 1" N (78)
2 2ω r

and the damped circular natural frequency of each mode is given by:

ωdr = ω r 1 − ζ 2r (79)

Finally, the frequency response functions for the multiple degree of freedom system are
defined as:

(Φ jr )(Φ kr )
B jk (Ω ) = ∑
N
(80)
r =1 k r − Ω 2 m r + iΩc r
Complex Notation
Consider a function of the form:

x (t ) = Xe iΩt (1)

where X is complex valued:

X = a + bi (2a)

This can also be written as:

X = X e iθ (2b)

where X = a 2 + b 2 and θ = tan −1 (b a ) .

Substituting Eq. 2b into 1 we obtain:

x (t ) = X e i (Ωt +θ ) (3a)

which may be written as:

x (t ) = X (cos(Ωt + θ) + i sin (Ωt + θ)) (3b)

Taking the real part of Eq. 3b yields:

Re(x (t )) = X cos(Ωt + θ) (4a)

or the imaginary part yields:

Im(x (t )) = X sin (Ωt + θ) (4a)

Thus, using x (t ) = Xe iΩt and taking either the real or imaginary component is equivalent
to using either x (t ) = X cos(Ωt + θ) or x (t ) = X sin (Ωt + θ) , respectively. The primary
advantage of using the exponential form is that derivatives with respect to time are
conveniently manipulated.

S-ar putea să vă placă și