You are on page 1of 16

Journal of Sound and Vibration (1992) lSa( 17, 1-16

MODAL ANALYSIS OF ROTATING SHAFTS:


A BODY-FIXED AXIS FORMULATION APPROACH

R. P. S. HAN AND J. W.-Z. Zu


Department of Mechanical Engineering, University of Manitoba, Winnipeg,
Manitoba, Canada R3 T 2h72

(Received 21 May 1990, and injinalform 10 April 1991)

A spinning Timoshenko beam subjected to a constant moving load is analyzed using a


modal expansion technique. The formulation is based on a body-fixed axis reference system,
and thus should be applicable to handle any general cross-sections. Using this alternative
approach, dynamic quantities such as natural frequencies, mode shapes and system
responses are easily computable. It is shown that simply supported spinning Timoshenko
beams possess two pairs of natural frequencies corresponding to each mode shape. Through
appropriate simplifications, the coupled eighth order differential equations governing the
problem are degenerated into a set of uncoupled fourth order equations. With this simplified
theory, closed form expressions for natural frequencies and system transient response are
derived. However, one pair of the natural frequencies is lost in the process, but this is
acceptable as the frequencies of this pair are much higher than those of the retained pair.
In addition, a linearized approximation of the natural frequency, similar to the one given
for Euler-Bernoulli beams, is also proposed for a Timoshenko shaft. Numerical simulations
are performed to demonstrate the characteristics of the response.

1. INTRODUCTION
Recently, there has been once again a tremendous interest in the understanding of the
dynamics of spinning structures. This is attributed to the importance of such knowledge
in an expanding area of application, which ranges from rotors and satellite booms for the
aerospace industries to machining operations in the manufacturing sector. In the past few
decades, there have been numerous papers in the domain of rotor dynamics. However, the
vast majority of these investigations are limited to Euler-Bernoulli beam model analysis
and subjected to static loadings. An excellent review of these studies was documented by
Dimentberg [l] and updated by Eshleman [2]. For moving loads on non-spinning struc-
tures, such as vehicular traffic on a bridge, a very comprehensive treatise was published
by Fryba [3]. In this paper, only the most recent progress in the dynamics of spinning
structures employing varying degrees of sophistication in beam theories, ranging from
Euler-Bernoulli to Rayleigh and Timoshenko beam models, is reported. Of particular
interest is the dynamic response of these spinning structures to traveling loads. Our motiva-
tion for this research is to obtain an accurate dynamic analysis of a spinning workpiece
as it is being cut by a moving tool traversing at a constant speed in a machining operation.
Since most workpieces are short and thick, a Timoshenko beam model is most appropriate
for the structural modelling.
A scan of the recent literature reveals that both analytical and numerical solutions have
been attempted. In the area of analytical techniques, Bauer [4] was one of the first to give
a very thorough treatment of spinning structures. However, his analysis is restricted to the
Euler-Bernoulli beam model. One important finding of his work is that, for an Euler-

0022-460X/92/130001 + 16 $03.00/O fQ 1992 Academic Press Limited


2 R. P. S. HAN AND J. W.-Z. ZU

Bernoulli beam, the natural frequency changes linearly with the spin rate when defined in
a body-fixed reference frame. Filipich et al. [5] extended Bauer’s method to analyze the
free vibrations of an elastically restrained, spinning Euler-Bernoulli beam, taking into
account the orthotropic properties of the supports. Huang and Chen [6] presented further
results using Bauer’s technique, but for spinning asymmetric beams loaded by moving
harmonic forces. Rotating shafts, modeled as Rayleigh beams, were analyzed by Lee and
Jei [7] and Lee et al. [8], with the latter presenting solutions for moving loads. Their
analysis also included the effects of gyroscopic or Coriolis acceleration terms. In a sub-
sequent paper, Katz et al. [9] analyzed a spinning Timoshenko beam loaded by moving
forces. A finite integral transform technique was employed to obtain the transient response.
Numerical methods are also popular in solving spinning dynamics. Hashish and Sankar
[IO] developed a finite element model to solve a spinning Timoshenko beam under the
influence of non-moving stochastic loads. Leung and Fung [ 111, also using a finite element
procedure, presented natural frequency results for general spinning space structures based
on the Euler-Bernoulli beam theory.
In this paper, the problem of a spinning Timoshenko beam subjected to moving loads
as used in Katz et al. [93 is re-solved using a modal analysis technique. Unlike their finite
integral transform procedure, this alternative method gives a better physical understanding
to the problem. The solution not only consists of the system transient response, but equally
important, eigenquantities such as natural frequencies and mode shapes. It is found that
a simply supported, spinning Timoshenko beam has two pairs of natural frequencies,
corresponding to the forward and backward precessions, for each mode shape. This is
quite different from that of a spinning Euler-Bernoulli beam of any boundary conditions
(Bauer [4]) and a simply supported, spinning Rayleigh beam (Lee et al. [8]), which possess
only one pair of natural frequencies. Using Hamilton’s principle, the equations of motion
are derived. These differential equations are coupled eighth order equations and, through
appropriate simplifications, they reduce to a set of uncoupled fourth order equations. It
is found that the convergence rate of the simplified analysis is comparable to the exact
eighth order analysis. However, one pair of natural frequencies vanishes in the process,
but this is acceptable since the frequencies of the lost pair are much higher compared to
those of the retained pair. Closed form expressions of the retained frequencies and the
system transient response are derived. In addition, linearized approximations of the natural
frequencies for a Timoshenko beam are also proposed. The resulting expressions are useful
at low spin rates and for low mode shapes. The analysis, as presented, has also been
specialized to handle a spinning Euler-Bernoulli shaft.

2. EQUATIONS OF MOTION
Figure 1 shows a beam of length 1, spinning along its longitudinal axis at a constant
rate 52. A set of inertial co-ordinates, OX YZ, with the origin 0 at the left end of the beam,
is adopted. In addition to this inertial frame, a rotating co-ordinate system is also defined.
This body-fixed co-ordinate system is denoted by Oxyz. Note that in the undeformed
configuration, the inertial co-ordinate 02 coincides with the body-fixed co-ordinate Oz.
Let { and q represent the two centerline elastic displacements of the beam along the x and
y directions, respectively. Taking into account the shear deformations and assuming small
deformations, the bending angles a< and a, in the two perpendicular planes of motion xz
and yz are given by

=z+g,,
ag
ar7
a,=z+9r12
az
(1)
ROTATING SHAFTS
r

Figure I. Spinning shaft subjected to moving loads.

in which 4~ and 4, are the corresponding shear angles, computed from VS/~AG and V,J
KAG respectively. The kinetic and strain energies for a beam of toss-sectional area A and
moment of inertia Z are

T=; ‘pA[~2+rj2+02(~2+q2)+2R(&j-[n)]dz
s0

+; ‘pl[d;+ai;-S22(a~+a;)] dz, (2)


s0

I
S=:EZ '(apfaf)dz+$cAG N5’-as)2+W-~,)21 dz, (3)
s0 s0

where p, E, G and K are the mass density, Young’s modulus, shear modulus and shear
coefficient, respectively. Also, spatial derivatives d( )/dz are indicated by ( )‘, while time
derivatives d( )/dt are denoted by ( ‘). The spinning beam is loaded by moving transverse
forces, Px(Z, t) and Py(Z, t), which remain parallel to the inertial co-ordinates OX and
0 Y. Each of these forces has a constant magnitude and moves with a constant velocity in
the OZ direction. Transforming these inertial system based forces into the body-fixed axis
based forces Pe(z, t) and P,(z, t) gives

(4)

where the transformation matrix is given by

cos nt sin i2t


[Tl= (5)
-sin Qt cosf2t I ’

The work expression now becomes

(6)
4 R. P. S. HAN AND J. W.-Z. ZU

Applying Hamilton’s principle, the equations of motion for a spinning Timoshenko


beam, subjected to travelling forces in a body-fixed co-ordinate formulation, are

(9)

It should also be pointed out that the formulation, as given here, is based on a body-
fixed axis frame and hence the mass moment of inertia remains constant at all times. This
implies that this proposed analysis can handle any cross-sections. In contrast, the inertial
formulation given by Han and Zu [ 121 is only applicable to cross-sections possessing axial
symmetry, such as a circular section. Thus, in this respect, a body-fixed formulation is
more superior.

3. METHOD OF SOLUTION

The method of modal analysis is used analytically to solve the equations of a spinning
Timoshenko beam loaded by moving forces, namely those depicted by equations (7)-( 10).
Unlike the finite integral transform technique given by Katz et al. [9], this alternative
modal approach gives a better insight into the problem. In addition to obtaining the system
transient response, the method also yields eigenquantities such as natural frequencies and
mode shapes. However, as in reference [9], the analysis given here is applicable only to
simply supported boundary conditions. For other homogeneous boundary conditions, the
analysis is much more complicated, as the determination of these eigenquantities now has
to satisfy supplementary conditions, in addition to the boundary conditions. This area of
research is currently being investigated.
Introducing dimensionless variables P,, = nlr and { = Z/f, the solutions to the differential
equations of motion take the form

5 = “E, L(t) sin (Pm0, 17= f Mt) sin (LC), (11, 12)
?I=1

ae = f aen cos (M), a,= $ a,,(t) cos (L&C). (13, 14)


?l=l n=l
Note that the above assumed solutions automatically satisfy the simply supported boun-
dary conditions. Substituting equations (1 l)-( 14) into equations (7)-( 10) and invoking
orthogonality yields

(15)
ROTATING SHAFTS 5

(17)

(18)

where

q&l=2
s ’APCsin (Pn 0
op
dt, q,,=2

Solving for aen and a,,,, from equations (15) and (16) yields
s
‘P
2
0 PA
sin (Pnc) d[. (19720)

at”- -~tC-zni.-ra'r.4i.)+~~~, (21)


PI
aw =-(~.+2n&d22rjn-q,.)+~ qn. (22)
KG/h

Substituting equations (21) and (22) into equations (17) and (18) results in the coupled
equations
k,~~-2Rk,ij:,+k2~~-2Rk,li.+kL5,=kSq5.+k,ij5”, (23)
k, ~~+2nk,~“ik,ij,+2Rk35’,+k,rl,=kSq,,+k,~,,, (24)

in which

k,=l+

(25)

k,=l+

Observe that the terms on the left side of equations (23) and (24) represent the unknown
elastic displacements, while on the right side are the prescribed moving loads. These two
fourth order equations are coupled by the spin rate L?. When the shaft is stationary,
these two equations become uncoupled, as expected. Also, unlike the equations of motion
formulated in an inertia co-ordinate system presented by Han and Zu [ 121, here there is
no coupling in the forcing terms. However, some of the ki coefficients in equation (25) are
more complicated in this approach as they now contain the 0’ term.

3.1. EXACT SOLUTION


Note that the two coupled equations in equations (23) and (24) represent an eighth
order system of partial differential equations. Solution of this system constitutes the exact
analysis, as no simplifications are assumed. Both the natural frequencies and system tran-
sient response are computed analytically by solving the coupled eighth order equations.
6 R. P. S. HAN AND J. W.-Z. ZlJ

For convenience of mathematical manipulations, the following parameters are defined

51 =<?I, r71=17n, 12 = k2/h,

52=&v tl2’ fin, 13 = b/h >


(26)
53=4;1, q3=?n, 14 = k4/k,,

54=4”n, 94= iim 15=k5/k,.

In view of equation (26), equations (23) and (24) can thus be compactly rewritten as

125)
= [CIW + IQ>, (27)
where
0 1 0 0 0 0 0 0
0 0 1 0 0 0 0 0
0 0 0 1 0 0 0 0
-14 0 -12 0 0 -2n1, 0 21R
[Cl= 0 0 0 0 0 1 0 0
, (28)
0 0 0 0 0 0 1 0
0 0 0 0 0 0 0 1
0 -201, 0 -2fl -14 0 -12 0

W=G 52 53 64 tll tl2 f73 ‘/4>*, (29)


IQ>= (0 0 0 Q<n 0 0 0 Qrp>*, (30)
and

(31~32)
The solution of the equations of motion in equation (27) comprises two parts. The first is
the homogeneous solution from which the free vibration characteristics are obtained, and
the second is the particular solution which, when combined with the homogeneous solu-
tion, yields the system response. To perform the free vibration analysis, let A represent the
eigenvalues of the matrix [Cl. For non-trivial solutions, the characteristic equation is
][c]--nq=o. (33)
The presence of the non-proportional gyroscopic or Coriolis acceleration term in the
equations of motion dictates that the eigenvalues appear as pure imaginary conjugate
pairs, while the eigenvectors ri, Fi appear as complex conjugate pairs. That is:
eigenvalues eigenvectors

A, = -A2 = io,, , {h>, {h),


&=-A4=i~n2, {r2>, {f22>, (34)
ilS=-&=io,3, {r3>, {F3;>,

&=-A8=iio,4, {r4>, {f4),

where O,i is the nth mode natural frequency of a spinning Timoshenko beam formulated
in a body-fixed axis system. Note that, unlike the spinning Euler-Bernoulli beams and
simply supported spinning Rayleigh beams, which have only two eigenvalues correspond-
ing to each eigenfunction, in the case of a simply supported spinning Timoshenko beam,
ROTATING SHAFTS 7
there arefiur eigenvalues, with two of them occurring at a higher order of magnitude.
Physically, these eigenvalues would represent the forward and backward whirl frequencies
of the shaft. The homogeneous solution can thus be written as

{X/J = WC1
{yc>+ PSI{Y,}, (35)
where

[~d=~(Gi k2h (r3)m (r4)9i 692 (r2h (r3h (r4)3], (36)

[Rl=[(rtk~ (r2)3 (r3),7 (r4)z (rh (r2)m (r3h (r4)9t], (37)


and

{Y,>=(A, coscn,t A2 cos a2t A3 cos o3t A4 cos co4t

B, cos wit B2 cos co2t B3 cos w3t B4 cos co&, (38)

{ Ys} =(-A, sin colt -A2 sin co2t -A3 sin m3t -A4 sin m4t
Bl sin w,t B2 sin 02t B3 sin w3t B4 sin m4t)T, (39)
where (ri)R and (ri)J are the real and imaginary parts of (ri) respectively. The arbitrary
constants AI-A4 and BI-B4 are evaluated from the assigned initial conditions. To evaluate
the particular solution, a constant magnitude force P in the OX inertial co-ordinate
direction and travelling at a constant speed o is prescribed:

(40)

where 6 is the Dirac delta function. Note that these forces are expressed in the inertial
frame and they must be transformed into the body-fixed axis system, in accordance with
equation (4): that is,

{~}=~Tl{~}={-‘pc~~~~~~~~~~_~~}. (41)

Substituting equation (41) into the loading function given by equations (19) and (20)
yields

P
q6n=-(sinIRIt-sin02t), q,,=~(cos12,t-cosi22t), (42,43)
PA PA

in which

and equations (31) and (32) become

&=: [(IS-.R:) sin 0, t-(Is--f2:) sin l2,t], (4)

Q,=; [(k42:) cos R, t-(1~42:) cosL12t]. (45)


8 R. P. S. HAN AND J. W.-Z. ZU

In view of the structure of equations (44) and (45), the particular solutions for en and
q, in equations (23) and (24) can be assumed to take the form

5, = C,, sin 0, t + Cc, sin L& t, q,=C,,c0s52~t+C*~C0Si2~f,

where the coefficients C,, , G2, C,, and C,,, are given by

1,-n:
cti=c&&
PA n~-l~R~+l~-2R(~~-l~Ri) I

i= 1,2. (46)

Hence the particular solution of equation (27), namely {X,,}, is

(X,}=(Ci sinIRIt+C2sin0zt, Ci51, cosR1t+CzR~cosR~t,


-C,52~sin~,t-C~R~sinR~t,-C,R~cosR,t-CzR~cosR~t,
C1cosRIt+Czcos~~t, -C,R, sinort-C2LJ2sinRzt,
-~IO:cosR1t-CzR:cosn,t,C,n:sinn,t+C~n:sin~,t)T. (47)

Applying superposition, the total solution of equation (27) is

IX> = PC1{yc>+ PSI{rs>+ (4J. (48)


Initial conditions are necessary for the evaluation of the arbitrary constants AI-A4 and
B,-B4. They are given by

5(O, 0 = 0, &O, 0 = 0,
tl(O, 0 = 0, MO, 0 = 0,
(49)
a#, 4-)= 0, d<(O, 0 = 0,
a ,(O, r) = 0, d ,(O, 0 = 0,
which, upon conversion so that they are suitable for use by equation (48), become

51=5n=o, 91= qn=O,


52=5’n=o, 92= 7i,=o,
(50)
&=4’,=0, t/3= ii,=O,
&=&=2P/(pA) q4= &=o.

The computed elastic displacements 5 and n must now be converted back to the inertial
co-ordinate system, and this is achieved by using the transformation matrix listed in
equation (5) ; that is,

{ ;}=VIT{ f }, (51)

where U and V are elastic displacements measured in the inertial co-ordinates OX and
OY.

3.2. SIMPLIFIED SOLUTION


It is observed that for most engineering materials, the coefficient k, in equation (25) is
very small, of the order I/(AG), compared with the other coefficients. Hence, neglecting
ki , the original coupled eighth order differential equations degenerate into simpler fourth
ROTATING SHAFTS 9

order equations; that is,

kz~~-21Rk,rj,+k,5,=k,q;,, kz ijn+ 2flk, 5’,,+ k‘, tin= k5 qr,n. (52953)


These equations can easily be uncoupled as
k; $, + (2kz kz,+ 4S2“k:)& + ki &, = Q5n, (54)
k: & + (2kz kq + 4R *k:) fjn+ k: q,, = et,,, , (55)
where

= k2 k5 &,, + 252kJ k5 (i,,n+ kdks q<n,


Qt,, (56)
Q,,,,= k2 ks c&n- 2f2kj k5 4<n+ k4k5 qrln. (57)
The nth mode natural frequencies @,I and w,,* can now be obtained in closed form, and
are

@nl,nZ = -; ~(k2k4+2R2k:)=F~(kzk~+29*k:)*-k:k:. (58)


2

Note that the original four frequencies predicted by the eighth order theory are now
reduced to two in this fourth order model. It should be pointed out that instability occurs
when the condition (k2 kq + 2l2 *k:)* = (kz kq + 20 *k:)* -k: ki is satisfied. The system
response is computed from equations (1 l)-( 14), and the results are
m
{(c, t) = 1 {At1 cos wnl t + BgI sin wnl t + A,, cos co,,*t + Bt2 sin w,:! t
n=l
+ C, sin IR, t + C2 sin a2 t} sin ( /L c), (59)

77(1, t)= : {A Visin o,,~ t + B,, cos 61,~t + Aq2 sin w,,*t + B+? cos con2t
ll=I
+C,cos0,t+C2cosf12t}sin(j?,<), (60)
in which the eight constants, A,, , B,, , . . . , A 02, B,,*, are evaluated using initial conditions.
These constants are not completely independent of each other as they are related by
equations (52) and (53) :

b&d, +k,M,, =2flk3wz,A,,, C-k, d, + WB,, = -2% @,I 41,


(61)
C-k:! 42 + kdAc2 = 2flkj ~“2 A 112, (-k2 co:2 + k4)Be2 = -252kJ a,,2 B,,2.

The other two remaining constants, C, and C2, are given by

c ,p -kzkgG:-2l2k~k&, +k,k,
’ pA k:R:-(2k2k,+4R2k:)R:+k:’
(62)
P k2ksIR:+252k3k512,-k,k,
c*=-
pA k:l2;-(2k~k,+4l2*k:)G!:+k:’

Using the transformation matrix given by equation (5), these solutions can easily be
transformed back to the inertial frame so that their corresponding physical responses can
be realized. The validity of these simplified solutions will be examined later with the aid
of numerical simulations. In the next section, some results pertaining to Euler-Bernoulli
beams are presented.
10 R. P. S. HAN AND J. W.-Z. ZU

4. EULER-BERNOULLI BEAM SOLUTIONS


It is of interest to present here the response of an Euler-Bernoulli beam subjected to
moving loads, with the formulation derived in a body-fixed co-ordinate system. This is
easily done, simply by retrieving the Euler-Bernoulli solutions from the Timoshenko results
given earlier. It should be pointed out that these equations of motion, when derived in an
inertial co-ordinate system, are identical to the non-spinning Euler-Bernoulli beam sub-
jected to moving loads. Such solutions are already well studied, with analytical results
available in Fryba [3]. However, this inertial-based formulation is valid only for cross-
sections possessing axial symmetry such as a circular section. For more general cross-
sections, a body-fixed formulation such as presented here must be used. Putting k, =O,
kZ = k3 = ks = 1, and k4 = EIp i/(pA14) - l2 * in the Timoshenko model represented by equa-
tions (23) and (24) (or by equations (52) and (53)) the differential equations for the
Euler-Bernoulli model become

(63)

(64)

where qtn and q,,,, are defined in the same way as before. It should be observed that these
equations of motion formulated,in the body-fixed axis are coupled by the spin rate 0.
These coupled equations can easily be uncoupled, or alternatively, can be deduced from
equations (54) and (55) with the appropriate simplifications inserted. Since the solution
process follows the same pattern as the procedure given for the simplified solutions, the
details will not be repeated here. Only the solutions will be shown. Thus, the nth mode
natural frequencies for an Euler-Bernoulli beam are

(65)

where o. = P f Jm is the “at-rest” frequency when the shaft is stationary. Note that
these natural frequencies results are identical to those obtained by Bauer [4]. The system
response is given by

HC, VB= f {A , cos w,“,“t+ Bi sin wE:t + A2 cos c&f? + B2 sin w,“,“t
n=l
+ C, sin L!, t + C2 sin a2 t} sin (/In c), (66)

rl(C,OEB=
f {A, sin o,“,“t- BI cos &? - A2 sin mi?Bt+ B2 cos w:f?
II=1
+C,cosR,t+C2cosR2t}sin(p,[), (67)
where the constants AI-B2 are determined from initial conditions and the coefficients C,
and C,, computed using equation (62), are suitably adjusted for use in the Euler-Bernoulli
beam model. As before, these solutions can easily be transformed back to the inertial co-
ordinate system for computing the physical responses.

5. NUMERICAL RESULTS AND DISCUSSION


Some examples were solved to examine the validity of the simplified solutions and
to study the response characteristics. For the convenience of performing the numerical
ROTATING SHAFTS 11
simulations, the following dimensionless variables were defined: a = u/o,,, , the non-dimen-
sional load speed; iii,, = Ln/w,,~, the non-dimensional rotational speed; o= U/u,, the non-
dimensional deflection along the X-direction ; v= V/u,,, the non-dimensional deflection
along the Y-direction ; f= {/ uo, the non-dimensional deflection along the x-direction ; fj =
q/uo, the non-dimensional deflection along the y-direction; and p = ma/l, the Raleigh
beam coefficient; where u,, is the critical speed of the load, given by o,,= (lr/,)Jm,
CD&is the at-rest nth-mode natural frequency, and u. = P1’/48EI is the static deflection of
a simply supported beam. The Rayleigh beam coefficient /I provides a measure of the
slenderness effects of a shaft. For a circular section, it relates the shaft diameter to its
length.
To perform the numerical simulations, the basic data employed is the same as used by
Katz et al. [9] ; namely,
a=O*ll-1.5, K =0.9,
4 =2.5, p = 7700 kg/m3,
p=o.15, E = 207 GPa,

I= 1 m, G = 77.6 GPa.
The variation of natural frequencies with the rotational speed parameter fi is sketched
in Figure 2. These results are computed for the case of p =O* 15, and correspond to the
first three mode shapes. Note that both the quantities on the x and y axes have been
normalized with respect to the lowest at-rest nth-mode natural frequency wno. As discussed,
there are two pairs of natural frequencies for each mode shape in Timoshenko beams.
These two pairs reduce to one in using the simplified analysis, the results of which are also
plotted for comparison with the full solutions. Excellent agreement is obtained, especially
at low spin rates and for lower mode shapes.
The linearity of the plot suggests that further simplification of equation (58) is possible.
Recall that, for an Euler-Bernoulli beam, the frequency relationship is linear and thus
very convenient to use. In a similar fashion, if a linearized form of equation (58), which
is derived for the Timoshenko beam, can be obtained with good accuracy, then the resulting
formula will also be very valuable. For most engineering materials at low spin rates, the
coefficients k2 and k3 in equation (25) are approximately related by
k2zk3zk<<w;, (68)

where k is given by
k=l+(piEZ/12~,4G) (69)
and o. is the at-rest frequency for Euler-Bernoulli beams defined earlier. Introducing these
approximations into equation (58) the following linear relationship results:

~,,,,2=l~=W. (70)
Observe that in the limit of an Euler-Bernoulli beam, that is for k = 1, equation (70)
reverts back to equation (65). The results of the linearized approximation are also shown
in Figure 2. The agreement for the first mode shape is particularly good.
Next, the system transient response is presented. The X-deflection 0 computed at the
location of the load, for a varying load speed parameter a, is shown in Figure 3(a).
Observe that the peaks of each plot shift to the right side of the shaft with increasing
speed. This is expected, since the applied load is moving from left to right. For the selected
values of fi = 2.5, and p = 0.15, the largest u deflection for the a = O-5 curve is 171% of
12 R. P. s. HAN AND J. W.-z. zu

(c)
7-

6- 4th

55

Figure 2. Variation of natural frequencies with spin-rate for p =O, 15; (a) first mode, (b) second mode and
(c) third mode. --, Full solution; - - -, simplified solution; ., linear approximation.

o- I 0

I.2
0.05
30
; 0.9 2
3
I3 ,< 0.00
0.6

-0.05

-0. IO I I
0.0 0.2 0.4 0.6 0.0 I.0

Figure 3. Deflections measured in an inertial reference frame for fi = 2.5 and p = 0.15; (a) ii deflection and
(b) V deflection. -, Full solution; . . ., simplified solution.
ROTATING SHAFTS 13
the static deflection due to a stationary load of the same magnitude. The Y-deflection v
is shown in Figure 3(b), and its presence, albeit small, indicates that an axially moving
force traveling on a spinning shaft not only produces deflections in the OXZ inertial plane,
but also in the OYZ plane, which is perpendicular to the plane of loading. For this case,
the largest amplification in the v deflection for the a = 0.5 curve is 12.8% of the statically
loaded case. It should be mentioned that these out-of-plane deformations are not observed
in spinning Euler-Bernoulli beams. This is not surprising, as the equations of motion for
spinning Euler-Bernoulli beams, when formulated in an inertial frame, resemble those of
non-spinning beams, and the resulting system response will therefore not exhibit such out-
of-plane deformations.
The results of the simplified analysis are also depicted in Figure 3. The agreement with
the exact analysis is reasonably good for the i7 deflection, but poor for the v deflection.
The reason becomes obvious when these same physical responses are plotted in the body-
fixed frame as shown in Figure 4. Observe that in this body-fixed co-ordinate system, the
agreement obtained is excellent. Errors are introduced when these responses are trans-
formed to the inertial reference frame. It should also be pointed out that the solutions of
the exact analysis coincide identically with the results of Katz et al. [9], which were obtained
using a Laplace transform technique.

20

I5

I ,o

O-5

00

-0 5

-I 0

-I .5
00 0.2 0.4 0.6 0.8 I.0 o-o 02 04 0.6 0.8 IO
[=2//

Figure 4. Deflections measured in a body-fixed reference frame for fi = 2.5 and p =O, I5 ; (a) 5 deflection and
(b) rj deflection. -, Full solution; ., simplified solution.

The variation of the maximum U- and J’-deflections with the Rayleigh beam coefficient
p is given in Figure 5. Since /? is the slenderness ratio, this plot is useful as it quantifies
the influence of the effects due to shear deformation and rotary inertia on spinning struc-
tures. Unlike the previously given deflection quantities, which were computed at the loca-
tion of the load, the maximum deflections as presented here are obtained by scaning for
the largest deflection for each load position and thus should reflect the true maximum
value. Note that in Figure 5(a), the U-deflections has been normalized with respect to the
maximum deflections predicted by the Euler-Bernoulli beam theory. However, for the V-
deflection, in Figure 5(b), it is not possible to normalize the result in this manner, as the
deflection predicted by the Euler-Bernoulli beam theory in this direction is zero. Hence,
the normalization is carried out with respect to u. . In Figure 5(a) it is shown that in order
to keep modelling errors under 5%, the Euler-Bernoulli beam model should be replaced
by the Timoshenko beam model for values of p 20.13. The V-deflections in Figure 5(b)
are not considered in choosing this /I criterion, since they are of an order of magnitude
smaller than the U-deflections.
14 R. P. S. HAN AND J. W.-Z. ZU

0.0 0.1 0.2 0.3 0.4


-0.51
0.0 0.1 0.2 0.3 0.4
B
Figure 5. Variation of maximum deflections with Rayleigh beam coefficient for fi =2.5; (a) CJ,,,/U~~C and
(b) V&Juo. -, E-B beam; . . ., ~=~.~~;___,~.5;-._.-,~.~; -___, 1.5.

The variation of the maximum U-deflections with the load speed a for different spin
rates D is shown in Figure 6(a). As illustrated, the influence of R is negligible. This is not
unexpected, as it is a well-known fact that quantities measured in an inertial reference
frame do not exhibit a strong dependency on J2 [ 121. As a matter of fact, in the limiting
case of an Euler-Bernoulli beam, its equations of motion do not contain any 0 terms at
all when formulated in an inertial reference frame. Observe that the maximum V-deflections
given in Figure 6(b) apparently depict some dependency on Sz. This is due to the fact that
the Y-axis has been enlarged for clarity and, also, because when the shaft is non-rotating,
the V-deflection is zero.
In all of the cases studied, the convergence criteria is defined by prescribing a given
tolerance. This generally corresponds to a summation of lo-15 terms.

2.5
(a) (b)

2-c

S?
L I.5
1
3

IC

0.5
0.0 0.5 I.0 I.5 2.0 0.0 0.5 IO I.5 2.0
cY=VI!+
Figure 6. Variation of maximum deflections with load speed parameter for p =0.15; (a) V,,,/uO and (b)
V,,,/~~.-,R=2.5;---,2.0;....,~.5;_.--,~.0; _.._.._, 0.5;_ ._._, 0.0.

6. SUMMARY AND CONCLUSIONS


A simply supported spinning Timoshenko beam subjected to an axially moving load is
solved analytically using a modal expansion technique. The theory is formulated in a body-
fixed reference frame and thus has the advantage of being able to handle any general cross-
sections. Both eigenquantities such as natural frequencies and mode shapes, and the system
ROTATING SHAFTS 15
transient response are easily computable using this approach. The governing equations are
coupled eighth order differential equations which can be reduced, through appropriate
simplifications, to a set of uncoupled fourth order equations. From this simplified analysis,
closed form expressions for natural frequencies and the system response are presented.
From the results of the numerical simulations, the following conclusions can be reached.
(a) A linearized approximation of the natural frequencies for spinning Timoshenko beams
was proposed. The agreement with the exact solutions is very good for lower mode shapes
and at low spin rates. (b) The system response shows that an axially moving load on a
spinning Timoshenko shaft produces deflections not only in the plane of loading, but also
in the plane perpendicular to the loading. (c) It is recommended that the Euler-Bernoulli
beam model be replaced by the Timoshenko beam model for values of /3 2 0.13 and fi > 1.
This will ensure that the errors in modelling will not exceed 5%. (d) The influence of spin
rates on quantities measured in an inertial reference frame is negligible. This is clearly
demonstrated by the plot of the maximum deflections versus load speed for varying R.

ACKNOWLEDGMENTS
This work was supported by an operating grant from the Natural Sciences and Engineer-
ing Research Council of Canada.

REFERENCES
1. F. M. DIMENTBERG 1961 Flexural Vibrations of Rotating Shafts. London: Butterworth.
2. R. L. ESHLEMAN 1978 Flexible Rotor Bearing Systems Dynamics. New York: ASME
Publications.
3. L. FRYBA 1972 Vibration of Solids and Structures under Moving Loads. Groningen, The
Netherlands : Noordhoff.
4. H. F. BAUER 1980 Journal of Sound and Vibration 72, 177-189. Vibration of a rotating uniform
beam, part I: orientation in the axis of rotation.
5. C. P. FILIPICH, M. J. MAURIZI and M. B. ROSALES 1987 Journal of Sound and Vibration 116,
475-482. Free vibrations of a spinning uniform beam with ends elastically restrained against
rotation.
6. S. C. HUANG and J. S. CHEN 1990 Journal of the Chinese Society of Mechanical Engineers 11,
63-73. Dynamic response of spinning orthotropic beams subjected to moving harmonic forces.
7. C. W. LEE and Y. G. JEI 1988 Journal of Sound and Vibration 126,345-361. Modal analysis of
continuous rotor-bearing system.
8. C. W. LEE, R. KATZ, A. G. ULSOY and R. A. SCOTT 1988 Journal of Sound and Vibration 122,
119-130. Modal analysis of a distributed parameter rotating shaft.
9. R. KATZ, C. W. LEE, A. G. ULSOY and R. A. Scorr 1988 Journal of Sound and Vibration 122,
131-148. The dynamic response of a rotating shaft subject to a moving load.
10. E. HASHISH and T. S. SANKAR 1984 Journal of Vibrations, Acoustics, Stress, and Reliability in
Design 106, 80-89. Finite element and modal analyses of rotor-bearing systems under stochastic
loading conditions.
11. A. Y. T. LEUNG and T. C. FUNG 1988 Journal of Sound and Vibration 125, 523-537. Spinning
finite elements.
12. R. P. S. HAN and J. W. Z. Zu 1992 Journal of the Franklin Institute (submitted). Analytical
dynamics of a spinning Timoshenko beam subjected to a moving load.

APPENDIX : NOMENCLATURE
A cross-sectional area
E Young’s modulus
G shear modulus
transverse moment of inertia of a cross-section
: length of the beam
16 R. P. S. HAN AND J. W.-Z. ZU

mode number
magnitude of moving load
external forces in the X and Y directions
,=,~;;?$;eo;~ra$i;&
normalized transverse deflection in the X direction
= P13/48EI, static displacement
speed of the moving load
= (n/l) Jm, critical speed of the load
deflection in the Y direction
normalized transverse displacement in the Y direction
= v/vCV,load speed parameter
bending angle in the x-z plane
bending angle in the y-z plane
= ma/l, Rayleigh beam coefficient
Dirac delta function
shear coefficient
eigenvalues
displacement in the x direction
mass density
displacement in the y direction
shear angle in the x-z plane
shear angle in the y-z plane
ith natural frequency of the nth mode
ith natural frequency of the nth mode for the Euler-Bernoulli beam
spin rate
=fl*(P,lO
= f2/o,, non-dimensional rotational speed
=d( )/dz, spatial derivative
= d( )/dt, time derivative

You might also like