Lagrangians For Damped Linear Multi-Degree-of-Freedom Systems

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Firdaus E.

Udwadia
Professor
Departments of Aerospace and
Mechanical Engineering,
Lagrangians for Damped Linear
Civil Engineering,
Mathematics, Multi-Degree-of-Freedom
Systems Architecture Engineering,
and Information and
Operations Management,
Systems
University of Southern California,
430 K Olin Hall, This paper deals with finding Lagrangians for damped, linear multi-degree-of-freedom
Los Angeles, CA 90089 systems. New results for such systems are obtained using extensions of the results for
single and two degree-of-freedom systems. The solution to the inverse problem for an
Hancheol Cho n-degree-of-freedom linear gyroscopic system is obtained as a special case. Multi-
Graduate Student degree-of-freedom systems that commonly arise in linear vibration theory with symmetric
Department of Aerospace and mass, damping, and stiffness matrices are similarly handled in a simple manner. Conser-
Mechanical Engineering, vation laws for these damped multi-degree-of-freedom systems are found using the
University of Southern California, Lagrangians obtained and several examples are provided. [DOI: 10.1115/1.4023019]
Los Angeles, CA 90089

1 Introduction provided some nonconservative systems to which their approach


can be applied.
The inverse problem for Lagrangian dynamics—also known as
In the present paper, the findings obtained in Refs. [2] and [3]
the inverse problem of the calculus of variations—is to obtain, for
are extended to dissipative, constant coefficient, linear MDOF
a system described by a given set of differential equations, a
systems. The emphasis is on obtaining the Lagrangians for these
Lagrangian function such that the corresponding Euler–Lagrange
MDOF systems in a simple manner, using insights obtained from
equations obtained using the calculus of variations yield the given
our understanding of the inverse problem for the SDOF and 2-
set of equations that describe the system. This problem has
DOF systems. It is shown that the solution to a gyroscopically
attracted many researchers in various fields of study for its useful-
damped linear system is easily found as a special case of the line-
ness; for example, in numerical mathematics a Lagrangian func-
arly damped case. Lagrangians for special linearly damped
tion provides us with approximate solutions to nonlinear ordinary
MDOF systems with symmetric stiffness and damping matrices
differential equations and in quantum physics, theories are based
are also obtained along with the corresponding Jacobi integrals,
on the Hamiltonian for the system. Bolza [1] gave a general pro-
which are conserved over time.
cedure for finding a Lagrangian for a single-degree-of-freedom
dissipative system. This was followed by Leitmann [2] who pro-
vided some examples of nonpotential forces and the correspond-
ing Lagrangians for which a variational principle exists. This
method was extended by Udwadia et al. [3] who used a more sys- 2 Lagrangians for Damped Linear
tematic derivation to obtain several classes of nonpotential forces Two-Degree-of-Freedom Systems
which could be used to obtain the equations of motion via varia- We begin with the problem of finding a Lagrangian function for
tional calculus. Recently, the semi-inverse method [4] has been a linear mass-spring-damper system in a single degree of freedom
considered due to its simplicity and applicability to many cases. whose governing equation of motion is given by
However, Refs. [1–4] consider only single-degree-of-freedom
(SDOF) systems and the analysis of this case is relatively easy x þ 2bx_ þ kx ¼ 0; m > 0; b; k  0
m€ (1)
because, in the nineteenth century, Darboux [5] proved that a
Lagrangian can always be found for the inverse problem for such
where xðtÞ is a generalized displacement of the mass, the dot
SDOF systems. For multi-degree-of-freedom (MDOF) systems,
denotes the differentiation with respect to time t, and m, b, and k
the configuration variables are coupled with one another and this
are the mass, damping, and stiffness coefficients, respectively,
makes it difficult to solve the inverse problem. The general condi-
which are all assumed to be constants. Unfortunately, Eq. (1)
tions for the existence of Lagrangians were apparently first
cannot be directly derived as the Euler–Lagrange equation from a
obtained by Helmholtz [6,7] and are usually referred to as the
variational principle because it does not satisfy the Helmholtz
Helmholtz conditions. Later, Douglas [8] analyzed in great detail
conditions [6,7] (see Eqs. (11)–(14)). However, in Refs. [2,3,11] it
the case of two degrees of freedom and obtained the necessary
is shown that the following Lagrangian function results in Eq. (1)
and sufficient conditions for their existence without utilizing these
conditions. Using Douglas’ results, Hojman and Ramos [9] pro-  
1 1
posed a simpler method to determine the existence of a Lagran- L ¼ eð2b=mÞt mx_2  kx2 (2)
gian for two-dimensional problems in which the potential function 2 2
does not explicitly contain the generalized velocities. Mestdag
et al. [10] derived the conditions under which there exists a More precisely, substituting Eq. (2) into the Euler–Lagrange equa-
Lagrangian and a dissipation function on the right hand side of the tion of the standard form ðd=dtÞð@L=@ xÞ_  ð@L=@xÞ ¼ 0 yields
more general form of the Euler–Lagrange equation. They also
eð2b=mÞt ½m€
xðtÞ þ 2bx_ðtÞ þ kxðtÞ ¼ 0 (3)
Manuscript received March 1, 2012; final manuscript received October 27, 2012;
accepted manuscript posted November 19, 2012; published online May 23, 2013. Since the exponential factor in Eq. (3) is always positive in time,
Assoc. Editor: Wei-Chau Xie. we can say that Eq. (3) is ‘equivalent’ to Eq. (1). This exponential

Journal of Applied Mechanics Copyright V


C 2013 by ASME JULY 2013, Vol. 80 / 041023-1

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 09/13/2013 Terms of Use: http://asme.org/terms


factor, however, plays a significant role, because Eq. (3) does sat-
isfy the Helmholtz conditions.
Thinking of the Lagrangian given by Eq. (2) as representing a
physical under-damped oscillator consisting of a ‘mass’ með2b=mÞt
and ‘potential energy’ ð1=2Þeð2b=mÞt kx2 , we find that the energy
of this oscillator does not decay with time as does that of the
under-damped system described by Eq. (1) [11]. In fact, its
energy, averaged over one period of the oscillator, remains a con-
stant and, hence, the Lagrangian given by Eq. (2) is not a physical
Lagrangian, but just a convenient mathematical tool to use
the machinery of the calculus of variations to obtain Eq. (3). In a Fig. 1 Linear 2-DOF mass-spring-damper system with b 5 2 m
similar manner, were the Lagangian in Eq. (2) to represent
an over-damped oscillator with ‘mass’ með2b=mÞt and ‘potential
energy’ ð1=2Þeð2b=mÞt kx2 , its energy would, in general, exponen-
tially increase with time, contrary to the behavior of the over-
a11 f 1 ¼ a11 ðx€1 þ a1 x_1 þ b1 x_ 2 þ c1 x1 þ d1 x2 Þ ¼ 0 (9a)
damped oscillator described by Eq. (1). It follows, then, that this
Lagrangian given in Eq. (2) does not describe the physical linearly
damped system, although the two systems are described by the a22 f 2 ¼ a22 ðx€2 þ a2 x_1 þ b2 x_ 2 þ c2 x1 þ d2 x2 Þ ¼ 0 (9b)
same equation of motion, namely, Eq. (1). However, a mathemati-
cal (not physical) Lagrangian may yet be useful in areas such as where aii ði ¼ 1; 2Þ are nonzero functions of t, x1 , x2 , x_ 1 , and x_ 2 ,
the development of approximate solutions of differential equa- namely, a11 ¼ a11 ðt; x1 ; x2 ; x_1 ; x_2 Þ and a22 ¼ a22 ðt; x1 ; x2 ; x_ 1 ; x_ 2 Þ.
tions and various numerical techniques. Next, we define the functions bi ði ¼ 1; 2Þ by
Taking a hint from Eq. (2), we next consider a two-degree-of-
freedom mass-spring-damper system using the Lagrangian given
by b1 :¼ a11 ða1 x_ 1 þ b1 x_ 2 þ c1 x1 þ d1 x2 Þ (10a)

 b2 :¼ a22 ða2 x_ 1 þ b2 x_ 2 þ c2 x1 þ d2 x2 Þ (10b)


1  1 
L ¼ ect m1 x_ 21  k1 x21 þ m2 x_22  k2 x22 þ b1 x_ 1 x2
2 2
 Now let us consider the n differential equations aij ðt; q; q_ Þ€ qj
1 2 n T
þ b2 x1 x_ 2 þ dx1 x2 (4) þ bi ðt; q; q_ Þ ¼ 0 ði; j ¼ 1; 2; …; nÞ, where q ¼ ½ q q    q  is
a generalized displacement n-vector that describes the motion of a
mechanical system in an n-dimensional configuration space. The
where mi , ki , bi , i ¼ 1; 2, and c and d are constants. Then the superscript “T” is used to denote the transpose of a vector (or a
corresponding Euler–Lagrange equations of motion are given matrix), and the summation convention is used for repeated
by indices. The question of whether such a system can be obtained
from a suitable Lagrangian Lðt; q; q_ Þ, through the use of
the Euler–Lagrange equations ðd=dtÞð@L=@ q_ i Þ  ð@L=@qi Þ ¼ 0,1
m1 x€1 þ cm1 x_1 þ ðb1  b2 Þx_ 2 þ k1 x1 þ ðb1 c  dÞx2 ¼ 0 (5a) appears to have been first investigated by Helmholtz [6,7]. The
necessary and sufficient conditions for the so-called ordered
m2 x€2 þ cm2 x_2 þ ðb2  b1 Þx_ 1 þ k2 x2 þ ðb2 c  dÞx1 ¼ 0 (5b) direct analytic representations are [7]

Depending upon the choice of the constants, Eqs. (5a) and (5b)
can represent various systems. For example, if we choose aij ¼ aji (11)
m1 ¼ m, m2 ¼ 2m, c ¼ 2, b1 ¼ b2 ¼ 0, k1 ¼ k2 ¼ 2k, and d ¼ k,
Eq. (4) becomes
@aij @aik
  ¼ (12)
1 2    @ q_ k @ q_ j
L ¼ e2t mx_1  2kx21 þ mx_22  kx22 þ kx1 x2 (6)
2
 
@bi @bj @ k @
and the corresponding equations of motion become þ ¼ 2 þ _
q aij (13)
@ q_ j @ q_ i @t @qk
x1 þ 2mx_1 þ 2kx1  kx2 ¼ 0
m€ (7a)
  
@bi @bj 1 @ k @ @bi @bj
x2 þ 4mx_2 þ 2kx2  kx1 ¼ 0
2m€ (7b)  ¼ þ _
q  (14)
@qj @qi 2 @t @qk @ q_ j @ q_ i
which describe a classically damped 2-DOF system. In fact,
Eqs. (7a) and (7b) describe the mechanical system shown in Fig. 1. where the summation convention is applied for repeated indices.
In order to obtain a more systematic approach to the inverse Equations (11)–(14) are a set of partial differential equations2 that
problem for a constant coefficient linear 2-DOF system we con- need to be satisfied by the 2n þ 1 independent variables t, q, and
sider the following equations of motion
1
Since we cannot differentiate the function Lðt; q; q_ Þ, with respect to a dependent
f 1 ¼ x€1 þ a1 x_1 þ b1 x_ 2 þ c1 x1 þ d1 x2 ¼ 0 (8a) variable, say q_ i , by @Lðt; q; qÞ=@
_ q_ i we mean @Lðt; s; rÞ=@ri js¼q , where t, s, and r are
f 2 ¼ x€2 þ a2 x_1 þ b2 x_ 2 þ c2 x1 þ d2 x2 ¼ 0 (8b) _
considered independent variables; similarly, by @Lðt; q; qÞ=@q i
r¼q_
, we mean
@Lðt; s; rÞ=@si js¼q .
where all coefficients are constant and we have divided each equa- 2
r¼q_
There is a slight abuse of notation here, since in Eqs. (11)–(14) the variables t, q,
tion by the corresponding masses m1 and m2 . More generally, we and q_ are considered to be independent, while in the equations of motion, x and x_
consider the following set of equations (see Eq. (8)) are considered to be functions of time t.

041023-2 / Vol. 80, JULY 2013 Transactions of the ASME

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 09/13/2013 Terms of Use: http://asme.org/terms


q_ everywhere in R2nþ1 . Throughout this paper we shall be dealing However, since a11 ¼ a11 ðt; x1 ; x2 Þ and a22 ¼ a22 ðt; x1 ; x2 Þ from
with Lagrangians that provide the so-called ordered direct analytic Eq. (25), taking the partial derivative of Eq. (26) with respect
representations of the equations of motion [7]. to the variable x_2 , we obtain @a11 =@x2 ¼ 0. In a similar fashion,
Applying these conditions to Eqs. (9a) and (9b) i; j; k ¼ 1; 2, taking partial derivatives with respect to x_ 1 yields @a11 =@x1 ¼ 0.
we therefore have q1 ¼ x1 , q2 ¼ x2 , and a12 ¼ a21 ¼ 0. Thus, By doing the same with Eq. (27), we obtain @a22 =@x1
Eq. (11) is automatically satisfied, and Eq. (12) yields ¼ @a22 =@x2 ¼ 0 so that

@a11 @a12 a11 ¼ a11 ðtÞ; a22 ¼ a22 ðtÞ (28)


¼ ¼0 (15)
@ x_ 2 @ x_1
@a21 @a22 and Eqs. (26) and (27) become
¼ ¼0 (16)
@ x_ 2 @ x_1 @a11
a11 a1 ¼ (29)
@t
from which we conclude that
and
a11 ¼ a11 ðt; x1 ; x2 ; x_1 Þ (17) @a22
a22 b2 ¼ (30)
a22 ¼ a22 ðt; x1 ; x2 ; x_2 Þ (18) @t

The general solutions to Eqs. (29) and (30) are


Next, Eq. (13) yields the following three equations

@a11 a11 ðtÞ ¼ c1 ea1 t (31)


ða1 x_ 1 þ b1 x_2 þ c1 x1 þ d1 x2 Þ þ a11 a1
@ x_1
a22 ðtÞ ¼ c2 eb2 t (32)
@a11 @a11 @a11
¼ þ x_ 1 þ x_2 ; ði ¼ 1; j ¼ 1Þ (19)
@t @x1 @x2 where c1 and c2 are nonzero constants. Substituting Eqs. (31) and
a11 b1 þ a22 a2 ¼ 0; ði ¼ 1; j ¼ 2 or i ¼ 2; j ¼ 1Þ (20) (32) into Eq. (24) yields
@a22 b1
ða2 x_ 1 þ b2 x_ 2 þ c2 x1 þ d2 x2 Þ þ a22 b2 c2 eb2 t ¼  c ea 1 t (33)
@ x_2 a2 1
@a22 @a22 @a22
¼ þ x_1 þ x_ 2 ; ði ¼ 2; j ¼ 2Þ (21)
@t @x1 @x2 Equation (33) should hold for all t  0, thus

Finally, Eq. (14) provides only one nontrivial equation for the a1 ¼ b2 (34)
case i ¼ 1; j ¼ 2, or i ¼ 2; j ¼ 1, which is b1
  c2 ¼  c (35)
@a11 1 a2 1
a1 x_ 1 þ b1 x_2 þ c1 x1 þ d1 x2 þ a11 d1
@x2 2
  and a11 and a22 become
@a22 1
 a2 x_1 þ b2 x_2 þ c2 x1 þ d2 x2  a22 c2
@x1 2 a11 ðtÞ ¼ c1 ea1 t (36)
 
1 @a11 @a11 @a22 @a22 b1
¼ b1 þ x_ 1 b1  a2  x_ 2 a2 (22) a22 ðtÞ ¼  c1 ea1 t (37)
2 @t @x1 @t @x2 a2

Thus, Eqs. (17)–(22) must be satisfied everywhere in R5 if Finally, we use Eq. (22) to obtain
_
Eqs. (9a) and (9b) can be obtained from a Lagrangian Lðt; x; xÞ,
where x ¼ ½ x1 x2 T , i.e., if a2 d1 þ b1 c2 ¼ a1 a2 b1 (38)
 
d @L @L In conclusion, we have the following possible case for the exis-
aij f j ¼  (23)
dt @ x_i @xi tence of a Lagrangian

We now consider two cases. b1


Case I: a11 ¼ c1 ea1 t ; a22 ¼  c ea 1 t ; a 1 ¼ b 2 ;
a2 1
2.1 Case I: a2 6¼ 0. If a2 6¼ 0, from Eq. (20) we have a2 d1 þ b1 c2 ¼ a1 a2 b1 ; a2 6¼ 0; b1 6¼ 0; c1 6¼ 0 (39)

b1 It should noted that Eq. (39) is symmetric; that is, if we inter-


a22 ¼  a11 ; b1 6¼ 0 (24)
a2 change the coefficients of Eqs. (9a) and (9b) with one another,
Eq. (39) still holds.
Then, from Eqs. (17) and (18)
2.2 Cases II and III: a2 5 0. If a2 ¼ 0, from Eq. (20) we
a11 ¼ a11 ðt; x1 ; x2 Þ; a22 ¼ a22 ðt; x1 ; x2 Þ (25) have
b1 ¼ 0 (40)
and Eqs. (19) and (21) become

@a11 @a11 @a11 Then, Eqs. (19) and (21) become


a11 a1 ¼ þ x_ 1 þ x_2 (26)
@t @x1 @x2
@a11 @a11 @a11 @a11
þ x_1 þ x_ 2  ða1 x_1 þ c1 x1 þ d1 x2 Þ ¼ a1 a11
@a22 @a22 @a22 @t @x1 @x2 @ x_ 1
a22 b2 ¼ þ x_ 1 þ x_2 (27)
@t @x1 @x2 (41)

Journal of Applied Mechanics JULY 2013, Vol. 80 / 041023-3

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 09/13/2013 Terms of Use: http://asme.org/terms


@a22 @a22 @a22 @a22 d1 c1 ea1 t ¼ c2 c2 eb2 t (52)
þ x_1 þ x_ 2  ðb2 x_2 þ c2 x1 þ d2 x2 Þ ¼ b2 a22
@t @x1 @x2 @ x_ 2
(42) which is required to hold for all t  0, so that

and Eq. (22) is now d1


a1 ¼ b2 ; c2 ¼ c ; c2 6¼ 0; d1 6¼ 0 (53)
c2 1
@a11
ða1 x_ 1 þ c1 x1 þ d1 x2 Þ þ a11 d1
@x2 We thus can obtain a second possible case for the existence of a
@a22 Lagrangian
¼ ðb2 x_ 2 þ c2 x1 þ d2 x2 Þ þ a22 c2 (43)
@x1 d1
Case II: a11 ¼ c1 ea1 t ; a22 ¼ c ea1 t ; a2 ¼ 0; b1 ¼ 0;
c2 1
However, a11 ¼ a11 ðt; x1 ; x2 ; x_ 1 Þ and a22 ¼ a22 ðt; x1 ; x2 ; x_2 Þ
according to Eqs. (17) and (18). Hence, taking the partial deriva- a1 ¼ b2 ; c2 6¼ 0; d1 6¼ 0; c1 6¼ 0 (54)
tive with respect to x_2 on both sides of Eq. (41), we find that
@a11 =@x2 ¼ 0. Using a similar argument for Eq. (42), we find that Finally, if c2 ¼ d1 ¼ 0 in Eq. (47), then from Eqs. (8a) and (8b),
@a22 =@x1 ¼ 0, so that the equations of motion are now

a11 ¼ a11 ðt; x1 ; x_1 Þ; a22 ¼ a22 ðt; x2 ; x_ 2 Þ (44) f 1 ¼ x€1 þ a1 x_ 1 þ c1 x1 ¼ 0 (55a)

and Eqs. (41) and (42) become f 2 ¼ x€2 þ b2 x_ 2 þ d2 x2 ¼ 0 (55b)


@a11 @a11 @a11
þ x_1  ða1 x_ 1 þ c1 x1 þ d1 x2 Þ ¼ a1 a11 (45) and the system is not coupled anymore. Each of these uncoupled
@t @x1 @ x_1 systems has the form given in Eq. (1) for which a Lagrangian and
@a22 @a22 @a22 the corresponding aii ’s (i ¼ 1; 2) are well-known and given in Eq.
þ x_2  ðb2 x_ 2 þ c2 x1 þ d2 x2 Þ ¼ b2 a22 (46) (2). Thus, this case is summarized as
@t @x2 @ x_2

Having the conditions in Eq. (44), we can simplify Eq. (43) fur- Case III: a11 ¼ c1 ea1 t ; a22 ¼ c2 eb2 t ; a2 ¼ 0; b1 ¼ 0;
ther as c2 ¼ 0; d1 ¼ 0; c1 6¼ 0; c2 ¼ 6 0 (56)
a11 d1 ¼ a22 c2 (47)
and the 2-DOF system is uncoupled.
where a11 ¼ a11 ðt; x1 ; x_ 1 Þ and a22 ¼ a22 ðt; x2 ; x_ 2 Þ, from Eq. (44). Table 1 summarizes the three cases of Eqs. (39), (54), and (56).
Now let us consider two cases. Additionally, corresponding to each case, it includes one Lagran-
First, if c2 6¼ 0 and d1 6¼ 0, then the left hand side of Eq. (47) gian, which shall be obtained later. It is to be noted that there are
is a function of t, x1 , and x_ 1 , whereas the right hand side is a many other possible Lagrangians which are different from the
function of t, x2 , and x_2 . Hence, a11 and a22 can be functions ones given in Table 1.
only of time t, i.e., a11 ¼ a11 ðtÞ and a22 ¼ a22 ðtÞ and, accordingly, As a simple application, let us consider a general 2-DOF mass-
Eqs. (45) and (46) become spring-damper system, shown in Fig. 2. Its equations of motion
are easily obtained as
@a11
¼ a1 a11 (48) b~1 þ b~2 b~2 k1 þ k2 k2
@t x€1 þ x_ 1  x_ 2 þ x1  x2 ¼ 0 (57)
@a22 m1 m1 m1 m1
¼ b2 a22 (49)
@t
b~2 b~2 þ b~3 k2 k2 þ k3
whose general solutions are, respectively, x€2  x_1 þ x_ 2  x1 þ x2 ¼ 0 (58)
m2 m2 m2 m2
a11 ðtÞ ¼ c1 ea1 t (50)
and each equation has been normalized by each mass for the cor-
a22 ðtÞ ¼ c2 eb2 t (51) respondence with the form given in Eqs. (8a) and (8b). The aim is
to find the conditions under which a Lagrangian would exist for
where c1 and c2 are arbitrary nonzero integration constants. With this system. First, by applying Case I given in Eq. (39) or Table 1,
Eqs. (50) and (51), Eq. (47) now reads as we obtain the following conditions

Table 1 Three cases when a Lagrangian function of the system described by Eqs. (8a) and (8b) exists and a corresponding
Lagrangian for each case

Case Conditions Lagrangian


 
I b1 1 2 b1 2 c1 2 b1 c2 b1 d2 2
a11 ¼ c1 ea1 t , a22 ¼  c ea1 t , a1 ¼ b2 , L ¼ ea1 t x_  x_  x þ x1 x2 þ x þ b1 x_ 1 x2
a2 1 2 1 2a2 2 2 1 a2 2a2 2
a2 d1 þ b1 c2 ¼ a1 a2 b1 , a2 6¼ 0, b1 6¼ 0, c1 6¼ 0
 
II d1 1 2 d1 2 c1 2 d1 d2 2
a11 ¼ c1 ea1 t , a22 ¼ c1 ea1 t , a2 ¼ 0, b1 ¼ 0, L ¼ ea1 t x_1 þ x_2  x1  d1 x1 x2  x2
c2 2 2c2 2 2c2
a1 ¼ b2 , c2 6¼ 0, d1 6¼ 0, c1 6¼ 0
   
III a11 ¼ c1 ea1 t , a22 ¼ c2 eb2 t , a2 ¼ 0, b1 ¼ 0, 1 2 1 2 1 1
L ¼ ea1 t x_1  c1 x1 þ eb2 t x_22  d2 x22
c2 ¼ 0, d1 ¼ 0, c1 6¼ 0, c2 6¼ 0 (uncoupled) 2 2 2 2

041023-4 / Vol. 80, JULY 2013 Transactions of the ASME

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 09/13/2013 Terms of Use: http://asme.org/terms


b1 a1 t
 e ðx€2 þ a2 x_ 1 þ a1 x_ 2 þ c2 x1 þ d2 x2 Þ ¼ 0 (63b)
a2

where c1 ¼ 1 is used. We next search for the conditions under


which a Lagrangian Lðt; x; x_ Þ exists, such that the corresponding
Euler–Lagrange equations yield Eqs. (63a) and (63b), that is
 
d @L @L
 ¼ aij ðt; x; x_ Þ€
xj þ bi ðt; x; x_ Þ (64)
dt @ x_i @xi
Fig. 2 General 2-DOF mass-spring-damper system
Expanding the total time derivative, we have the following identi-
b~1 þ b~2 b~2 þ b~3 b~1 þ b~2 ties [7]
¼ ; k2 ¼ b~2 ; b~2 6¼ 0 (59)
m1 m2 2m1
@2L
¼ aij (65a)
Moreover, Eq. (59) can be further divided into two cases whether @ x_ i @ x_ j
the two masses m1 and m2 are equal or not
@2L @2L @L
~  x_j þ  ¼ bi (65b)
m1 m2 b3 b~1 b~1 þ b~2 @ x_i @xj @ x_i @t @xi
b~2 ¼  ; k2 ¼ b~2 ; b~2 6¼ 0;
m2  m1 m 2 m 1 2m1
m1 6¼ m2 ; and b~1 ; b~3 k1 ; k3 are arbitrary (60a) where i; j ¼ 1; 2. Knowing aij and bi from Eqs. (39), (63), and
(64), we obtain the following Lagrangian, using Eqs. (65a) and
(65b)
b~1 þ b~2 ~ b~3 þ b~2
b~1 ¼ b~3 ;k2 ¼ b~2 ¼ b2 ; b~2 ¼
6 0;  
2m 2m 1 b1 2
~ L ¼ ea1 t x_21  x_ 2 þ x_1 pðt; x1 ; x2 Þ þ x_2 qðt; x1 ; x2 Þ
m1 ¼ m2 ¼ m; and b2 ; k1 ; k3 are arbitrary (60b) 2 2a2
þ r ðt; x1 ; x2 Þ (66)
Next, we apply Case II, given by Eq. (54) or Table 1, to obtain the
conditions where pðt; x1 ; x2 Þ, qðt; x1 ; x2 Þ, and rðt; x1 ; x2 Þ are arbitrary func-
tions of their arguments within the requirements that
b~1 b~3
b~2 ¼ 0; ¼ ; k2 ¼
6 0; and
m1 m2
@p @q
m1 ; m2 ; k1 ; k3 are arbitrary (61)  ¼ b1 ea1 t ;
@x2 @x1
   
It is straightforward to check that the mechanical system shown in @p @r b 1 c2
 ¼ ea1 t c1 x1 þ a1 b1  x2 ;
Fig. 1 falls into this case. @t @x1 a2
Finally, we apply Case III, given by Eq. (56) (see Table 1), to
@r @q b1 a1 t
have the conditions  ¼ e ðc2 x1 þ d2 x2 Þ (67)
@x2 @t a2
b~2 ¼ 0; k2 ¼ 0; and m1 ; m2 ; b~1 ; b~3 ; k1 ; k3 are arbitrary
For example, if we choose
(62)
In this case, the system is uncoupled.
In brief, if a given 2-DOF mass-spring-damper system pðt; x1 ; x2 Þ ¼ b1 x2 ea1 t ; qðt; x1 ; x2 Þ ¼ 0;
described by Fig. 2 does not satisfy at least one condition given  
c1 b1 c2 b1 d2 2
by Eqs. (60)–(62), then it appears to be not possible to find a r ðt; x1 ; x2 Þ ¼ ea1 t  x21 þ x1 x2 þ x2 (68)
Lagrangian so that the Euler–Lagrange equation ðd=dtÞð@L=@ x_i Þ 2 a2 2a2
ð@L=@xi Þ ¼ 0 (i ¼ 1; 2) yields an ordered direct analytic repre-
sentation of the equation of motion of the system. In particular, then the Lagrangian in Eq. (66) becomes
from Eq. (60) we see that when the two masses m1 and m2 (see
Fig. 2) are connected solely by a damper, i.e., when b~2 6¼ 0 and  
k2 ¼ 0, there appears to be no Lagrangian Lðt; x1 ; x2 ; x_1 ; x_2 Þ that 1 2
a1 tb1 2 c1 2 b1 c2 b1 d2 2
L¼e x_  x_  x þ x1 x2 þ x þ b1 x_1 x2
provides an ordered direct analytic representation of the equations 2 1 2a2 2 2 1 a2 2a2 2
of motion of the system described by Eqs. (57) and (58). (69)
Until now we have derived the necessary and sufficient condi-
tions for which there exists a Lagrangian function of the 2-DOF
system given by Eqs. (8) and we have obtained three cases, which is shown in Table 1.
Eqs. (39), (54), and (56), summarized in Table 1. We next address For Case II, following the same procedure shown in the previ-
the question of finding a Lagrangian for each of these cases. Since ous Case I, we can again obtain Lagrangian functions and one
a Lagrangian for Case III is already known in Refs. [2,3,11] (also possible Lagrangian is
in Eq. (2)), we focus on obtaining a Lagrangian function for Cases
I and II.  
1 2 d1 2
First, for Case I, from Eq. (39), we know that the equations of L ¼ ea1 t x_1 þ x_ 2 þ x_1 uðt; x1 ; x2 Þ þ x_2 vðt; x1 ; x2 Þ
2 2c2
motion are given by
þ wðt; x1 ; x2 Þ (70)
   
b 1 c2
e a1 t
x€1 þ a1 x_1 þ b1 x_ 2 þ c1 x1 þ a1 b1  x2 ¼ 0 (63a) where uðt; x1 ; x2 Þ, vðt; x1 ; x2 Þ, and wðt; x1 ; x2 Þ are arbitrary func-
a2 tions of their arguments within the requirements that

Journal of Applied Mechanics JULY 2013, Vol. 80 / 041023-5

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 09/13/2013 Terms of Use: http://asme.org/terms


@u @v @u @w where M and K are symmetric matrices and B is a skew-
¼ ;  ¼ ea1 t ðc1 x1 þ d1 x2 Þ; symmetric matrix. We note that since c ¼ 0, the Lagrangian is a
@x2 @x1 @t @x1
  physical Lagrangian (see the remarks that follow Eq. (3)).
@v @w d1 d2
 ¼ ea1 t d1 x1 þ x2 (71) As previously stated, Eq. (77) has a skew-symmetric damping
@t @x2 c2 matrix B. However, in many practical applications and, especially
in the theory of linear vibrations, the equations of motion have
For example, if we choose symmetric stiffness and damping matrices. In order to encompass
such systems, we choose c 6¼ 0 and BT ¼ B. Then, with the
uðt; x1 ; x2 Þ ¼ 0; vðt; x1 ; x2 Þ ¼ 0; Lagrangian given in Eq. (75), the Euler–Lagrange equation yields
 
c1 2 d1 d2 2
wðt; x1 ; x2 Þ ¼ ea1 t x1 þ d1 x1 x2 þ x2 (72) M€x þ ðcM þ 2BÞx_ þ ðcB þ KÞx ¼ 0 (79)
2 2c2
When the skew-symmetric matrix B ¼ 0, Eq. (79) reduces to the
then the Lagrangian Eq. (70) becomes
equation
 
1 d1 2 c1 2 d1 d2 2
L ¼ ea1 t x_ 21 þ x_ 2  x1  d1 x1 x2  x2 (73) M€x þ cMx_ þ Kx ¼ 0 (80)
2 2c2 2 2c2
which describes a proportionally damped system. The Lagrangian
which is listed in Table 1. Comparing Eqs. (7a) and (7b), which from which this equation is obtainable is simply given, using
are the equations of motion of the system shown in Fig. 1, with Eq. (75), by
Case II, we see that
 
ct 1 T 1 T
k L¼e x_ Mx_  x Kx (81)
a2 ¼ 0; b1 ¼ 0; a1 ¼ b2 ¼ 2; c2 ¼  ; 2 2
2m
k 2k k
d1 ¼  ; c1 ¼ 1; c1 ¼ ; d2 ¼ (74) for any matrix M > 0 and any symmetric matrix K. Having dis-
m m m posed of the case B ¼ 0, from this point on we shall then concen-
trate on the case when the skew-symmetric matrix is B 6¼ 0.
and the Lagrangian given in Eq. (73) then reduces to the one given We would thus want the matrices cM þ 2B and cB þ K in
in Eq. (6). The obtained Lagrangians are summarized in Table 1. Eq. (79) to be symmetric, where Bð6¼ 0Þ is a skew-symmetric
matrix. The required conditions are not obvious, therefore, let us
3 Lagrangians for Special Constant Coefficient consider a 3-DOF system, which, by extension, will help us to
Multi-Degree-of-Freedom Linear Systems adduce the general procedure for handling linearly damped
MDOF systems. We start by considering diagonal mass matrices.
If the number of degrees of freedom of a system is much greater If we have
than two, then solving the Helmholtz conditions becomes quite
complex and, hence obtaining Lagrangians for ordered direct ana- 2 3 2 3
m1 0 0 k1 0 0
lytic representations of n-degree-of-freedom systems by solving 6 7 6 7
these conditions becomes, in general, extremely difficult, if not M¼4 0 m2 0 5; K¼40 k2 0 5;
nearly impossible. In this section, we therefore use ideas from 0 0 m3 0 0 k3
SDOF and 2-DOF systems to expand our thinking to MDOF sys- 2 3
tems and, hence, altogether bypass the need for solving the Helm- 0 b12 b13
holtz conditions. For two-degree-of-freedom systems, we used the 6 7
B ¼ 4 b12 0 b23 5 (82)
Lagrangian given by Eq. (4), which we now generalize for an
MDOF system as b13 b23 0
 
ct 1 T 1 T T then Eq. (79) becomes
L¼e x Mx  x Kx  x Bx
_ _ _ (75)
2 2
2 32 3 2 32 3
m1 0 0 x€1 cm1 2b12 2b13 x_1
where x ¼ ½ x1 x2    xn T is a generalized displacement n- 6 76 7 6 76 7
vector, and M and K are the n by n symmetric mass and stiffness 4 0 m2 0 54 x€2 5 þ 4 2b12 cm2 2b23 54 x_2 5
matrices, respectively. In addition, the scalar c and the n by n 0 0 m3 x€3 2b13 2b23 cm3 x_3
matrix B are determined according to the requirements of the 2 32 3 2 3
problem, as shall be shown later. Then the equation of motion k1 cb12 cb13 x1 0
obtained via the Euler–Lagrange equation is 6 76 7 6 7
þ 4 cb12 k2 cb23 54 x2 5 ¼ 4 0 5 (83)
   
M€x þ cM þ B  BT x_ þ K  cBT x ¼ 0 (76) cb13 cb23 k3 x3 0

First, if we choose c ¼ 0 and BT ¼ B (skew-symmetric), and the damping and stiffness matrices of this system are not sym-
Eq. (76) becomes metric. However, noting the negative sign in the term that
involves x_ 22 in Eq. (69), if we choose to use the same B matrix as
M€x þ 2Bx_ þ Kx ¼ 0 (77) in Eq. (82) and

which is the general equation of motion for an n-degree-of-free- 2 3 2 3


dom linear mechanical system with gyroscopic damping. The m1 0 0 k1 0 0
corresponding Lagrangian is M¼4 0 m2 0 5; K ¼ 4 0 k2 05 (84)
0 0 m3 0 0 k3
1 1 1 1
L ¼ x_ T Mx_  xT Kx  xT Bx_ ¼ x_ T Mx_  xT Kx þ x_ T Bx
2 2 2 2 in our Lagrangian given in Eq. (75), then the equations of motion
(78) that we obtain are

041023-6 / Vol. 80, JULY 2013 Transactions of the ASME

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 09/13/2013 Terms of Use: http://asme.org/terms


2 32 3 2 32 3
m1 0 0 x€1 cm1 2b12 2b13 x_1 that is, i ¼ 2 and j ¼ 3 in this case, then we should choose the ele-
6 76 7 6 76 7 ments of the B matrix by the rule:
4 0 m2 0 54 x€2 5 þ 4 2b12 cm2 2b23 54 x_2 5
0 0 m3 x€3 2b13 2b23 cm3 x_3 (1) b23 ¼ 0, and
2 32 3 2 3 (2) after we delete the 2nd and 3rd rows and columns, the
k1 cb12 cb13 x1 0 remaining elements should be set to zero, i.e., set b14 ¼ 0
6 76 7 6 7
þ 4 cb12 k2 cb23 54 x2 5 ¼ 4 0 5 (85)
In brief, we should choose the following B matrix
cb13 cb23 k3 x3 0 2 3
0 b12 b13 0
or, equivalently 6 b12 0 0 b24 7
B¼6 4 b13
7 (90)
2 32 3 2 32 3 0 0 b34 5
m1 0 0 x€1 cm1 2b12 2b13 x_1 0 b24 b34 0
6 76 7 6 76 7
4 0 m2 0 54 x€2 5 þ 4 2b12 cm2 2b23 54 x_2 5
0 0 m3 x€3 2b13 2b23 cm3 x_3 Now using the Lagrangian given by Eq. (75), with M and K
2 32 3 2 3 defined in Eq. (89) and B defined in Eq. (90), the equations of
k1 cb12 cb13 x1 0
6 76 7 6 7 motion given by Eq. (79) become
þ 4 cb12 k2 cb23 54 x2 5 ¼ 4 0 5 (86)
2 32 3
cb13 cb23 k3 x3 0 m1 0 0 0 x€1
6 0 m2 0 0 7 6 x€ 7
6 76 2 7
in which the damping and stiffness matrices are now both sym- 6 76 7
4 0 0 m3 0 54 x€3 5
metric if we set b13 ¼ 0.
We now generalize this observation to general n-DOF systems. 0 0 0 m4 x€4
2 32 3
Our aim is to find the appropriate matrices M, K, and B which, cm1 2b12 2b13 0 x_1
when inserted into the Lagrangian given in Eq. (75), will result in 6 2b cm 0 2b 76 x_ 7
6 12 2 24 76 2 7
the Euler–Lagrange equation given in Eq. (79) that is of the form þ6 76 7
4 2b13 0 cm3 2b34 54 x_3 5
xþB
M€  x_ þ Kx
 ¼0 (87) 0 2b24 2b34 cm4 x_4
2 32 3 2 3
where M  is a positive definite diagonal matrix and the matrices B  k1 cb12 cb13 0 x1 0
and K  are both symmetric matrices, such as in Eq. (86). 6 cb k2 0 cb24 76 x2 7 6 0 7
7 6 7 6
6 12 7
Consider the diagonal matrices M ¼ diagðm1 ; m2 ; …;mn Þ and þ6 76 7 ¼ 6 7 (91)
4 cb13 0 k3 cb34 54 x3 5 4 0 5
K ¼ diagðk1 ; k2 ; …; kn Þ, where n  2. We propose to: (i) change
the signs of some of the elements of these matrices, and (ii) pro- 0 cb24 cb34 k4 x4 0
vide a procedure to make the matrix M  positive definite and the
matrices B  and K symmetric. We do this in the following manner. or
If we place negative signs on mi , mj , mk , …, ki , kj , kk , … 2 32 3
(i; j; k; … ¼ 1; 2; …; n, and i < j < k <   ), then the elements of m1 0 0 0 x€1
the skew-symmetric matrix B, which is given by 6 0 0 76 7
6 m2 0 76 x€2 7
2 3 6 76 7
0 b12 b13    b1n 4 0 0 m3 0 54 x€3 5
6 b12 0 b23    b2n 7 0 0 0 m4 x€4
6 7 2 32 3
6 b13 b23 0    b3n 7
B¼6 7 (88) cm1 2b12 2b13 0 x_1
6 .. .. .. .. .. 7 6 2b
4 . . . . . 5 6 12 cm2 0 2b24 7 6 x_ 7
76 2 7
b1n b2n b3n    0 þ6 76 7
4 2b13 0 cm3 2b34 54 x_3 5
should have its elements altered by the following rule: 0 2b24 2b34 cm4 x_4
2 32 3 2 3
k1 cb12 cb13 0 x1 0
(1) the elements are set so that bij ¼ 0, bik ¼ 0, bjk ¼ 0, … and
6 cb k2 0 cb24 76 x2 7 6 0 7
76 7 6
(2) after deleting the ith, jth, kth, … rows and ith, jth, kth, … 6 12 7
þ6 76 7 ¼ 6 7 (92)
columns of the B matrix in Eq. (88), the remaining ele- 4 cb13 0 k3 cb34 54 x3 5 4 0 5
ments of B are set to zero
0 cb24 cb34 k4 x4 0
Clearly, if we want to place only one negative sign, say, on mi
and ki (i ¼ 1; 2; …; n), then only the second rule (2) applies, since and Eq. (92) has symmetric damping and stiffness matrices along
B is skew-symmetric. Additionally, changing the signs of all the with a positive definite mass matrix. The corresponding Lagran-
mi ’s and ki ’s (i ¼ 1; 2; …; n), i.e., m1 ; m2 ; …; mn and k1 ; k2 ; …; kn , gian is given by Eq. (75).
will result in B ¼ 0, which is a case already considered in Eqs. More generally, when placing negative signs on mi , mj , mk , …,
(80) and (81), although in a more general manner. ki , kj , kk , …, (and following the procedure described earlier) we
For example, in a 4-DOF system, if we have have 2n different choices, each of which yields the corresponding
2 3 2 3 equations of motion in the form of Eq. (87). However, by symme-
m1 0 0 0 k1 0 0 0
try, half of them are equivalent. We can extend this symmetry to
6 0 m 0 0 7 6 0 k 0 07
6 2 7 6 2 7 the situation where we consider placing no negative signs on any
M¼6 7; K ¼ 6 7;
4 0 0 m3 0 5 40 0 k3 0 5 of the mi ; ki ;i ¼ 1; 2; …; n (and following the procedure) to be
equivalent to placing negative signs on every mi ; ki ;i ¼ 1; 2; …; n,
0 0 0 m4 0 0 0 k4
2 3 which yields, as previously mentioned B ¼ 0. Due to this symme-
0 b12 b13 b14 try, we therefore have 2n1 differently ‘structured’ systems of
6 b 0 b23 b24 7 equations of motion of the form of Eq. (87) (each with differently
6 12 7
B¼6 7 (89) structured matrices B  and K)
 for which the corresponding Lagran-
4 b13 b23 0 b34 5
gians can be obtained by using Eq. (75). Furthermore, for each
b14 b24 b34 0 system structure one can use any (permissible) parameter values.

Journal of Applied Mechanics JULY 2013, Vol. 80 / 041023-7

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 09/13/2013 Terms of Use: http://asme.org/terms


Thus, for large n, one generates Lagrangians for numerous or- Yet, it can give the equations of motion of certain systems consist-
dered direct representations of different n-DOF damped linear ing of a chain of masses in which each mass is connected to its
systems. neighbors by linear dashpots and springs. The strength of the pro-
As a special case, in an n-DOF system (n  2), if one chooses cedure introduced in this section is that it totally bypasses the
2 3 Helmholtz conditions, which are near-impossible to solve for arbi-
m1 0 0 0  0 trarily large finite values of n.
6 7 To exemplify what has been discussed thus far, we consider a
6 0 m2 0 0  0 7
6 7 3-DOF system described by Eq. (86) with b13 ¼ 0, where M, B,
6 7
6 0 0 m3 0  0 7 and K are given by
6 7
M¼6 7; 2 3 2 3
6 0 0 0 m4    0 7
6 7 m1 0 0 0 b12 0
6 . . . . . . 7 6 7 6 7
6 .. .. .. .. .. .. 7 M ¼ 4 0 m2 0 5; B ¼ 4 b12 0 b23 5;
4 5
0 0 0 0    ð1Þnþ1 mn 0 0 m3 0 b23 0
2 3
2 3 k1 0 0
k1 0 0 0  0 6 7
6 7 K ¼ 4 0 k2 0 5 (95)
6 0 k2 0 0  0 7
6 7 0 0 k3
6 7
60 0 k3 0  0 7
6 7
K¼6 7; Using the matrices given in Eq. (95) in the Lagrangian function
60 0 0 k4    0 7
6 7 given by Eq. (75), the resulting Euler–Lagrange equations of
6 . . . . . . 7
6 .. .. .. .. .. .. 7 motion are given by Eq. (86), which is of the form of Eq. (87),
4 5
where
0 0 0 0  ð1Þnþ1 kn
2 3 2 3 2 3
0 b1 0 0  0 m1 0 0 cm1 2b12 0
6 7  ¼6
M 4 0 m2 0 5; B
7 ¼6 4 2b12 cm2
7
2b23 5;
6 b1 0 b2 0  07
6 7
6 7 0 0 m3 0 2b23 cm3
6 0 b2 0 b3  07 2 3
6 7 k1 cb12 0
B¼6 7 (93)
6 0 0 b3 0  07  6 7
6 7 K ¼ 4 cb12 k2 cb23 5 (96)
6 . .. .. .. .. .7
6 .. . . . . .. 7 0 cb23 k3
4 5
0 0 0 0  0
In the theory of linear vibrations, one often considers proportion-
ally damped systems. One can then particularize the damping
using the matrices given by Eq. (93) in the Lagrangian given  of this system to have proportional damping so that
matrix B
by Eq. (75), the Euler–Lagrange equation of motion given by
Eq. (79) yields  ¼ aM
 þ bK

B (97)
2 32 3
m1 0 0 0  0 x€1
6 0 m 76 x€ 7 where a and b are some constants. Substituting Eq. (96) into
6 2 0 0    0 76 2 7
6 76 7 Eq. (97), we have the following relationships
6 0 0 m 3 0    0 76 x€3 7
6 76 7
6 0 0 m4    0 7 6 7 2 ki 2
6 0 76 x€4 7 a¼c ; b¼ ; i ¼ 1; 2; 3 (98)
6 . .. .. .. . . . 767 7 c mi c
6 . 6 . 7
4 . . . . . .. 54 .. 5
0 0 0 0    mn x€n that is, for proportional damping to be possible, each ratio ki =mi
2 32 3 (see Eq. (95)) should be the same. It is straightforward to verify
cm1 2b1 0 0  0 x_ 1
that Eq. (98) still holds for general n-DOF (i ¼ 1; 2; …; n) tridiag-
6 2b1 cm2 2b2 0  0 76 x_ 2 7
76
6 7 onal systems whose equations of motion are described by
6 76 7 Eq. (94). Now, consider a mass-spring-damper system of the type
6 0 2b2 cm3 2b3    0 76 x_ 3 7
6 76 7 shown in Fig. 3 that commonly appears in the theory of linear
þ6 0 0 2b3 cm4    0 7 6 7
6 76 x_ 4 7 vibrations.
6 76 . 7
6 .. .. .. .. .. .. 76 . 7 The equations of motion of the system are
4 . . . . . . 54 . 5
0 0 0 0  cmn x_ n 2 323 2 32 3
2 32 3 2 3 1 0 0 x€1 2 1 0 x_1
k1 cb1 0 0  0 x1 0 6 76 7 6 76 7
6 cb m4 0 1 0 54 x€2 5 þ b4 1 2 1 54 x_2 5
6 1 k2 cb2 0  07 6x 7 607
76 7 6 7
2
6 76 7 6 7 0 0 1 x€3 0 1 2 x_3
6 0 cb2 k3 cb3  0 76 x3 7 6 0 7 2 32 3 2 3
6 76 7 6 7 2 1 0 x1 0
þ6 0 0 cb3 k4  07 6 7¼6 7
6 76 x4 7 6 0 7 6 76 7 6 7
6 76 . 7 6 . 7 þ k4 1 2 1 54 x2 5 ¼ 4 0 5 (99)
6 .. .. .. .. .. .. 76 . 7 6 . 7
4 . . . . . . 54 . 5 4 . 5 0 1 2 x3 0
0 0 0 0  kn xn 0
(94) with b 6¼ 0. (When b ¼ 0, the forces can be derived from a poten-
tial and a Lagrangian for the system can be easily found.) We note
which is the equation for an MDOF system with tridiagonal the tridiagonal form of the damping and stiffness matrices and
symmetric damping and stiffness matrices of the form of that in Eq. (99) describes a proportionally damped system with a ¼ 0
Eq. (87). It should be noted that because the Lagrangian has nega- and b ¼ b=k: For this system to be described by the Lagrangian
tive quantities in the mass matrix, it is not a physical Lagrangian. given in Eq. (75), in which the matrices M, B, and K are specified

041023-8 / Vol. 80, JULY 2013 Transactions of the ASME

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 09/13/2013 Terms of Use: http://asme.org/terms


Hence, the Jacobi integral is given by Eq. (103) as

1
c T
c   1 c
I¼ y_ þ y M y_  y  y_ T B þ BT y þ yT Ky  yT By
2 2 2 2 2
(106)

and since the matrix B is assumed to be skew-symmetric in this


paper, Eq. (106) simplifies to
Fig. 3 General 3-DOF mass-spring-damper system
1
c T
c 1
I¼ y_ þ y M y_  y þ yT Ky (107)
2 2 2 2
by Eq. (95), we require a and b to also satisfy the conditions given
in Eq. (98). Hence c ¼ 2k=b and noting that mi ¼ m and ki ¼ 2k Rewriting this Jacobi integral in terms of x and x_ by using
(i ¼ 1; 2; 3), we require Eq. (104), we obtain the conservation law
 
k b 1 1
¼ (100) Iðt; x; x_ Þ ¼ ect ðx_ þ cxÞT Mx_ þ xT Kx ¼ constant (108)
b m 2 2

The application of this result to n-degree-of-freedom tridiagonal Thus, for an MDOF linear system whose equation of motion is
systems easily follows by extension. In fact, the condition given by Eq. (79), it is guaranteed that the function Iðt; x; xÞ _ in
Eq. (100), along with c ¼ 2k=b, should also be met for general Eq. (108) is conserved at all times. For example, corresponding to
n-DOF tridiagonal systems of the type given in Eq. (99) to be pro- the Lagrangian given in Eq. (102) for the dissipative proportion-
portionally damped. ally damped 3-DOF system, shown in Fig. 3, with the parameters
To illustrate this result, we apply it to a proportionally damped given in Eqs. (101a) and (101b), we obtain the conservation law
system described by Eq. (99) with the parameters m ¼ 0:64 kg,
b ¼ 0:8 Ns=m, and k ¼ 1 N=m. Noting that Eq. (100) holds, we   
I ðt; x; x_ Þ ¼ e2:5t 0:32 x_21  x_ 22 þ x_ 23 þ 0:8ðx1 x_ 1  x2 x_2 þ x3 x_ 3 Þ
find  
þ x21  x22 þ x23 ¼ constant (109)
2k b
c¼ ¼ 2:5 s1 ; a ¼ 0 s1 ; b¼ ¼ 0:8 s;
b k
m1 ¼ m2 ¼ m3 ¼ m ¼ 0:64 kg (101a) 4 Conclusions
Here we have discussed extensions of the inverse problem of
b b the calculus of variations for nonpotential forces to multi-degree-
b12 ¼  ¼ 0:4 Ns=m; b23 ¼ ¼ 0:4 Ns=m;
2 2 of-freedom systems. For a two-degree-of-freedom linear system
k1 ¼ k2 ¼ k3 ¼ 2k ¼ 2 N=m (101b) with linear damping, the conditions for the existence of a Lagran-
gian are explicitly obtained by solving the Helmholtz conditions.
by comparing Eq. (99) with Eqs. (87) and (96). The Lagrangian Three general cases for when such Lagrangians are guaranteed to
for this dissipative 3-DOF system given by exist are obtained, depending on the parameter values of the
coupled linear systems. The Helmholtz conditions are near-
  impossible to solve for general n-degree-of-freedom systems and,
1 1
L ¼ ect x_ T Mx_  xT Kx  xT Bx_ although they are explicit, from a practical standpoint they
2 2
provide little assistance in solving the inverse problem for such
2:5t
  2    systems. By using and generalizing the results for single-degree-
¼e 0:32 x_ 1  x_22 þ x_23  x21  x22 þ x23
of-freedom systems, a simple procedure that does not require the

þ 0:4ðx1 x_ 2 þ x_ 2 x3  x_1 x2  x2 x_ 3 Þ (102) use of the Helmholtz conditions and that is easily extended to n-
degree-of-freedom linear systems is developed. We specifically
then yields the equations of motion given in Eq. (99) for the spe- include systems that commonly arise in the theory of linear
cific values of m, k, and b, chosen in this example. vibrations—systems with positive definite mass matrices and sym-
As the last application of the results obtained thus far, we metric stiffness and damping matrices. The method yields several
propose a Jacobi integral that is conserved at all times for the new Lagrangians for linear multi-degree-of-freedom systems.
types of linear multi-degree-of-freedom systems considered here Conservation laws for such dissipative MDOF systems are also
(see Eq. (79)). When the Lagrangian does not explicitly contain obtained by finding the corresponding Jacobi integrals. Although
time (and the actual velocity is a virtual velocity), i.e., when the approach employed herein is simple and easy to apply to other
@Lðt; q; q_ Þ=@t ¼ 0, the Jacobi integral I given by Ref. [12] examples, a more rigorous and systematic way for arriving at gen-
eral solutions of the inverse problem for multi-degree-of-freedom
@L systems needs much more work. At present, it remains an open
I :¼ q_ T L (103) problem. Our discussion herein is restricted to linear damping and
@ q_ more general forms of nonpotential forces will be considered in
future work.
is conserved. The Lagrangian in Eq. (75), however, explicitly con-
tains time, however, by using the transformation Acknowledgment
The authors appreciate the comments on a very early draft of
y ¼ xeðc=2Þt (104)
this paper by Professor G. Leitmann at UC Berkeley.
it becomes
References
[1] Bolza, O., 1931, Vorlesungen über Varationsrechnung, Koehler und Amelang,
1
c T
c 1
c Leipzig, Germany.
L¼ y_  y M y_  y  yT Ky  yT B y_  y (105) [2] Leitmann, G., 1963, “Some Remarks on Hamilton’s Principle,” ASME J. Appl.
2 2 2 2 2 Mech., 30, pp. 623–625.

Journal of Applied Mechanics JULY 2013, Vol. 80 / 041023-9

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 09/13/2013 Terms of Use: http://asme.org/terms


[3] Udwadia, F. E., Leitmann, G., and Cho, H., 2011, “Some Further Remarks on [8] Douglas, J., 1941, “Solution of the Inverse Problem of the Calculus of Var-
Hamilton’s Principle,” ASME J. Appl. Mech., 78, p. 011014. iations,” Trans. Am. Math. Soc., 50, pp. 71–128.
[4] He, J.-H., 2004, “Variational Principles for Some Nonlinear Partial Differential [9] Hojman, S. and Ramos, S., 1982, “Two-Dimensional s-Equivalent Lagrangians
Equations With Variable Coefficients,” Chaos, Solitons Fractals, 19, pp. 847–851. and Separability,” J. Phys. A, 15, pp. 3475–3480.
[5] Darboux, G., 1894, Leçons sur la Théorie Générale des Surfaces, Vol. 3, [10] Mestdag, T., Sarlet, W., and Crampin, M., 2011, “The Inverse Problem for
Gauthier-Villars, Paris. Lagrangian Systems With Certain Non-Conservative Forces,” Diff. Geom.
[6] Helmholtz, H. v., 1887, “Uber die physikalische Bedeutung des Princips der Applic., 29, 2011, pp. 55–72.
kleinsten Wirkung,” J. Reine Angew. Math., 100, pp. 137–166. [11] Ray, J. R., 1979, “Lagrangians and Systems They Describe—How Not to Treat
[7] Santilli, R. M., 1978, Foundations of Theoretical Mechanics I: The Inverse Dissipation in Quantum Mechanics,” Am. J. Phys., 47, pp. 626–629.
Problem in Newtonian Mechanics, Springer-Verlag, New York, pp. 110–111 [12] Pars, L. A., 1972, A Treatise on Analytical Dynamics, Ox Bow, Woodbridge,
and 131–132. CT, p. 82.

041023-10 / Vol. 80, JULY 2013 Transactions of the ASME

Downloaded From: http://appliedmechanics.asmedigitalcollection.asme.org/ on 09/13/2013 Terms of Use: http://asme.org/terms

You might also like