Applied Mathematical Modelling: D. Palaniappan

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Applied Mathematical Modelling 35 (2011) 5494–5499

Contents lists available at ScienceDirect

Applied Mathematical Modelling


journal homepage: www.elsevier.com/locate/apm

A general solution of equations of equilibrium in linear elasticity


D. Palaniappan
Department of Mathematics and Statistics, Texas A&M University – Corpus Christi, 6300 Ocean Drive, Corpus Christi, TX 78412-5825, USA

a r t i c l e i n f o a b s t r a c t

Article history: A general solution of equations of equilibrium in linear elasticity is presented in cylindrical
Received 26 August 2009 coordinates in terms of three harmonic functions describing an arbitrary displacement field.
Received in revised form 7 January 2011 The structure of this solution is similar to the general solution given by Love (Kelvin’s solu-
Accepted 31 January 2011
tion) in spherical coordinates. Galerkin vector representation of our solution leads to an
Available online 12 February 2011
integral connecting the harmonic functions. The connections to Papkovich–Neuber and
Muki’s general representations are also provided. Suitable choices of the harmonic func-
Keywords:
tions in our new representation yield general solutions for axisymmetric deformations
Linear elasticity
General solution
due to Love, Boussinesq and Michell. Some unbounded deformations induced by singular
Galerkin representation forces are tabulated in terms of the scalar harmonic functions to justify the simple nature
Papkovich–Neuber solution of our representation. Exact solution of the half-space boundary value problem is also pro-
Boussinesq–Michell solution vided to demonstrate the power of our approach. The stress components computed via our
solution are also listed (see the Appendix).
Ó 2011 Published by Elsevier Inc.

1. Introduction

The vector equation of equilibrium in linear elasticity with no body forces, also known as Navier equations, for the disc-
placement field is given by
1
r2 u þ rðr  uÞ ¼ 0; ð1Þ
1  2m
where u is the displacement vector and m is the Poisson ratio. We note from Eq. (1) that the divergence of the displacement
vector u is a scalar harmonic function which in turn implies that the displacement is a vector biharmonic function. Therefore,
there can be at most six scalar harmonic functions associated with the general solutions of the equations of equilibrium. The
Galerkin and Boussinesq general solution [1,2] of the equations of equilibrium involve six scalar harmonic functions assem-
bled in a certain fashion. However, the six harmonic functions are not independent. Another famous solution due to Papko-
vich–Neuber [3,4] has four scalar harmonic functions and has been successfully employed in several contexts to construct
solutions of boundary value problems in linear elasticity. It is also mentioned that the fourth harmonic function in Papko-
vich–Nueber representation can be omitted without loss of generality (see [3], for instance). A simple explanation for reduc-
ing the six harmonic functions to three goes as follows. The three displacement components, each component is biharmonic,
constitute six unknown scalar harmonic functions which are connected (only) by three equations. Therefore, one would expect
three independent harmonic functions in a complete general solution of the vector Navier equation. Love [5] presented a general
solution (known as Kelvin’s solution [6]) of the equations of equilibrium in terms of three harmonic functions in spherical
coordinates. Kelvin’s solution is particularly suited to solve boundary value problems involving spherical inclusions. Naghdhi
and Hsu [7] gave another representation of the general solution in terms of three stress functions. Below, we provide a new
type of general solution of (1) in cylindrical coordinates in terms of three scalar harmonic functions which form a set of basis

E-mail address: devanayagam.palaniappan@tamucc.edu

0307-904X/$ - see front matter Ó 2011 Published by Elsevier Inc.


doi:10.1016/j.apm.2011.01.041
D. Palaniappan / Applied Mathematical Modelling 35 (2011) 5494–5499 5495

functions for the Navier equations for further computations. Our representation, like the solution due to Love/Kelvin and
Naghdhi and Hsu, splits the biharmonic vector function for the displacement into three scalar harmonic functions and there-
fore can be conveniently used, in principle, to solve problems in the presence of boundaries for a given deformation field.
Specifically, the given boundary value problem for the displacement vector can be reduced – via our solution representation
– to solving simpler problems for the three scalar harmonic functions with appropriate boundary conditions. The proposed
solution can be effectively utilized to generate solutions of particular boundary value problems involving three-dimensional
body profiles including cylindrical domains and half spaces. To the best of our knowledge, the present solution representa-
tion does not seem to exist in the context of linear elasticity.

2. A general solution in cylindrical coordinates

In many practical circumstances, it is often required to compute the displacement field in the presence of cylindrical do-
mains. Here we present a general solution of the Navier equations suitable for solving boundary value problems involving
body profiles that can be expressed conveniently in cylindrical coordinates. A general solution for the displacement field
u in terms of three independent harmonic functions in cylindrical coordinates (q, /, z) is
 
1 @ 3  4m @ P
u¼ rw þ q rP þ ez þ r  ð c
c ez vÞ; ð2Þ
2ð1  mÞ @q 2ð1  mÞ @z
where P(q, /, z), w(q, /, z), and v(q, /, z) are scalar harmonic functions, that is, r2{P, w, v} = 0. It can be easily verified that the
solution given by (2) satisfies the equations of equilibrium (1). The general solution representation (2) has two parts, namely,
a divergence-free part containing the functions w and v and non-zero divergence part containing the harmonic function P.
The components of displacement in the radial, tangential and azimuthal directions (uq, u/, uz) are given by
  
1 @w @ @P 1 @v
uq ¼ þq þ ; ð3Þ
2ð1  mÞ @ q @q @q q @/
  
1 1 @w @ @P @v
u/ ¼ þq  ; ð4Þ
2ð1  mÞ q @/ @ q q@/ @q
   
1 @w @ @P 3  4m @ P
uz ¼ þq þ ; ð5Þ
2ð1  mÞ @z @ q @z 2ð1  mÞ @z
and the divergence of (2) yields

ð1  2mÞ @ 2 P
ru¼ : ð6Þ
1  m @z2
It should be mentioned that in the hydrodynamical context, a general solution for the incompressible velocity field similar to
(2) was given by Happel and Brenner [8] many years ago. Their solution was employed to solve a variety of boundary value
problems involving cylindrical domains in the so called Stokes flow regime. We note that in the limit of Poisson ratio m ! 12
(incompressible limit), our general solution reduces to the one given in [8].
By combining the harmonic functions of all orders for the functions one can write the solution in an infinite series form as
1 
X   
1 @ 3  4m @ Pk
u¼ rwk þ q rPk þ ez þ r  ð c
c ez vk Þ : ð7Þ
k¼1
2ð1  mÞ @q 2ð1  mÞ @z

Here wk, vk, and Pk are arbitrary cylindrical harmonic functions of order k in (q, /, z) coordinates. Suitable solutions of La-
place equation for cylindrical inclusions are of the form [8]
Z 1
fPn ; wn ; vn g ¼ cosðn/ þ fan ; bn ; cn g fPn ðkÞ; wðkÞ; vðkÞgIn ðkqÞ cosðkz þ fdk ; nk ; fk gÞdk; ð8Þ
0

where In(kq) is the cylindrical Bessel function of order n. The infinite series solution representation for the displacement filed
given in (7) has the same structure as that given by Love [5] (also known as Kelvin’s solution [6]). By choosing harmonic
functions of required order together with the prescribed conditions at the surface of a body profile, the boundary value prob-
lem can be solved for a given induced deformation. Some unbounded deformation fields in terms of our scalar functions are
provided in Table 1. The unbounded deformation fields for many point singularities (Kelvinet, rotlet and their gradients are
given in [6], see Section 2.1.4) in terms of Green’s functions.
We now provide connections of our representation for the displacement field to other famous solutions in linear elastic-
ity. To this end, we first write the Helmholtz decomposition form of the solution (2) as
 
1 @P
u¼ $ wþr þ ð1  2mÞdivðzPc
ez Þ þ Curl½CurlðzPc
ez Þ þ vc
ez : ð9Þ
2ð1  mÞ @r
The Helmholtz decomposition form yields remarkable features of the vector field as noted in hydrodynamics [9]. Here we
exploit this form to connect our solutions to others. We define Galerkin–Boussinesq vector as
5496 D. Palaniappan / Applied Mathematical Modelling 35 (2011) 5494–5499

Table 1
Expressions for some elastic deformation fields in terms of the harmonic functions P, w, v. Here D = 2(1  m)U and F1, F2, and F3 refer to the axis of those
singular forces and R21 ¼ x2 þ y2 þ ðz  cÞ2 .

Fields P0(q, /, z) w0(q, /, z) v0(q, /, z)


Uniform deformation 0 D(qcos/, qsin/, z) 0
Kelvinet (asymmetric)  8Fpl
1 R1
q cos /
12m R1
4pl q cos /
F1 R1
8ð1mÞpl q sin /
Rotlet (asymmetric) 0 2 ðzcÞ
 8Fpl F2 ðzcÞ
qR1 cos /  16ð1 mÞpl qR1 sin /
Rotlet (axisymmetric) 0 0 F3 1
8pl R1

Z   
@P
G¼ q  P þ w dz c
ez þ Curlvc
ez : ð10Þ
@q
Now the general solution in Galerkin form can be written as [1]

1  2m
u ¼ Curl Curl G  grad div G ð11Þ
2ð1  mÞ
Using (10) for the Galerkin vector G in (11), it can be shown (after the use of some vector identities) that our solution (2) is
equivalent to the Galerkin representation of the displacement (11). Observe that the Galerkin vector connecting our scalar
functions has an integral. The integrand in the Galerkin vector can be seen in our general solution if we re-write our solution
(2) in the form
 
1 @P @P
u¼ r wþq P þ2 ez þ r  ðc
c ez vÞ: ð12Þ
2ð1  mÞ @q @z
Eq. (12) provides an alternative form of our general solution given in (2).
Now the choice

@A @P
¼q  P þ w; ð13Þ
@z @q
in Eq. (10), gives the Galerkin vector

G ¼ Aðq; /; zÞc
ez þ Curlvc
ez ; ð14Þ
where

r4 A ¼ 0 and r2 v ¼ 0: ð15Þ
Substitution of (14) in (11) yields precisely the representation form given by Muki [10] in terms of a biharmonic function and
a harmonic function. This representation has been used in variety of problems involving half spaces and discs. The relation
between the harmonic function w and P and the biharmonic function A (U in Muki’s notation) is given in (13). The harmonic
function v is related to the harmonic function w in Muki’s notation.
Another most widely used differential representation is the famous Papkovich–Neuber (PN) representation in terms of a
vector harmonic function B and a scalar harmonic function U given by

1
u¼B rðr  B þ UÞ; ð16Þ
4ð1  mÞ
where r2B = r2U = 0. If we take

@P
B¼2 c
ez þ Curlvc
ez ; ð17Þ
@z 
@P
U¼2 Pr w ; ð18Þ
@r
in (16), we obtain (2) or equivalently (12). Thus our representation is connected to the popular solutions for the displace-
ment field in linear elasticity. Connections to other existing solutions can be established in a similar fashion (see Section
3 for axisymmetric solutions).
As mentioned in Section 1 the exact solution representation given in (2) (or (12)) can be employed to solve numerous
boundary value problems in three-dimensions. In practice, a typical vector boundary value problem can be deduced to solv-
ing equivalent subproblems for the three scalar harmonic functions P, w, and v and this is one of the main advantages of our
solution. Below we provide an example to illustrate the efficacy of our solution approach.
D. Palaniappan / Applied Mathematical Modelling 35 (2011) 5494–5499 5497

2.1. Half-space displacement boundary value problem

Consider the boundary value problem in the half-space z P 0 in which the displacement field is specified on the bound-
ary. Let the given displacement be

u ¼ U; on z ¼ 0; ð19Þ

where U(x, y) = hU1, U2, U3i is a given vector field. In order to obtain the displacement field u satisfying the boundary condition
(19), we rewrite (12) in the form
 
1 @ 1 @P 3  4m @ P
u¼ r wþr PP  zr þ ez þ r  ðc
c ez vÞ: ð20Þ
2ð1  mÞ @r 2ð1  mÞ @z 2ð1  mÞ @z

The boundary condition (19) suggests that one can choose w ¼ P  r @@rP and therefore the required solution can be generated
by solving the boundary value problems for the harmonic functions P and v. Using (19) in (20) one obtains
Z Z
1
v1 ¼ U 1 ðn; gÞ ln½ðy  gÞ þ Rdndg; ð21Þ
2p
Z Z
1
v2 ¼  U 2 ðn; gÞ ln½ðx  nÞ þ Rdndg; ð22Þ
2p
Z Z
ð1  mÞ U 3 ðn; gÞ
P¼ dndg; ð23Þ
pð3  4mÞ R
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where R ¼ ðx  nÞ2 þ ðy  gÞ2 þ z2 . Note that on the boundary of the half-space z = 0

@ v1 @ v2 3  4m @ P
¼ U1 ; ¼ U2 ; ¼ U3
@y @x 2ð1  mÞ @z

yielding u = U. Thus the solution for the half-space displacement boundary value problem is
3  4m @ P 1 @ P @ v1 @v
u¼ c
ez  zr þ ex  2 c
c ey ; ð24Þ
2ð1  mÞ @z 2ð1  mÞ @z @y @x
where v1, v2 and P are given in ()(21)–(23). Other boundary value problems with specified displacement and/or stress con-
ditions can be treated in a similar fashion.

3. General axisymmetric solutions of Love, Boussinesq and Michell

For the axisymmetric deformation, the displacement components depend on q and z and this implies that the azimuthal
component is zero. In the following we provide connection between the general solutions of the famous axisymmetric solu-
tions in linear elasticity and ours.
Love’s solution:
If we take
@L @P
 ¼wþq  P; ð25Þ
@z @q
@P
r2 L ¼ 2 ; ð26Þ
@z
in (3)–(5) and noting that all the quantities are independent of the angular coordinate /, we obtain
 
1 @ @L
uq ¼  ; ð27Þ
2ð1  mÞ @ q @z
 
1 @ @L
uz ¼ r2 L  ; ð28Þ
2ð1  mÞ @z @z
where L(q, z) is the well-known axisymmetric Love’s stress function [5,11,12] satisfying the axisymmetric biharmonic equa-
tion r4L(q, z) = 0. Eqs. (27) and (28) are precisely Love’s solution for the Navier equations in the axisymmetric case.
Boussinesq solution:
By taking
 
1 @P
 ðB0 þ zBz Þ ¼ wþq P ; ð29Þ
2 @q
1 @P
Bz ¼ ; ð30Þ
2 @z
5498 D. Palaniappan / Applied Mathematical Modelling 35 (2011) 5494–5499

in (3)–(5) we obtain

1 @
uq ¼  ðB0 þ zBz Þ; ð31Þ
4ð1  mÞ @ q
1 @
uz ¼ Bz  ðB0 þ zBz Þ; ð32Þ
4ð1  mÞ @z

where r2B0 = r2Bz = 0. This is the known representation for the axisymmetric deformation due to Boussinesq [2,12,13].
Michell’s solution: Another commonly used solution in the context of axisymmetric deformation is due to Michell [14] (see
also [12,13,15]). If we set
 
@M M @P
 þ ¼wþq  ð3  4mÞP; ð33Þ
@q q @q
M @P
r2 M  2 ¼ 2 ; ð34Þ
q @q
in (3)–(5), we obtain Michell’s solution given by [12–15]
 
M 1 @ @M M
uq ¼ r2 M   þ ; ð35Þ
q 2ð1  mÞ @ q @ q q
2
 
1 @ @M M
uz ¼ þ ; ð36Þ
2ð1  mÞ @z @ q q
where the Michell’s function M(q, z) satisfies ðr2  q12 Þ2 Mðq; zÞ ¼ 0 and we have used the result
   
@M M 1 @ M
r2 þ ¼ qr2 M 
@q q q @q q
in the derivation of the above solution.
The other forms of the axisymmetric solutions, including the general solution of Burgatti [12,16], can also be deduced
from our general solution (3)–(5) for the displacement.

4. Conclusion

It is shown that our new general solution representation of Navier equations of equilibrium contains the famous solutions
that have been widely discussed and used in the literature. The displacement field given in Eq. (2) or equivalently in (12) or
(20) in terms of three independent harmonic functions may be added to the exclusive list of general solution representations
in the theory of linear elasticity. The half-space displacement boundary value problem (illustrated in Section 2.1) exemplify
the power of our solution approach. In general, our solution representation can be employed to solve a variety of boundary
value problems, like deformation in cylindrical domains and sphere moving in a cylindrical domain, for a given displacement
field. Additional related boundary value problems including the question of completeness of the general solution (2) will be
discussed elsewhere.

Acknowledgement

Major portion of this work was done while the author was at the Department of Mathematics, Texas A&M University,
College Station, TX-77843 and the author wishes to thank Professor Al Boggess for his help.

Appendix A

The components of stress in terms of the functions P, w, and v, computed using (3)–(5), are given by (in cylindrical coor-
dinates [17])

rqq 1 1 1
¼ ðw þ Pqq þ qPqqq þ 2mPzz Þ þ vq/  2 v/ ; ð37Þ
2l 2ð1  mÞ qq q q
 
r// 1 1 1 1 1 1 1
¼ w // þ wq þ P q//  P// þ Pqq þ 2 m Pzz  v/q þ 2 v/ ; ð38Þ
2l 2ð1  mÞ q2 q q q2 q q
rzz 1
¼ ðw þ qPqzz þ ð3  2mÞPzz Þ; ð39Þ
2l 2ð1  mÞ zz
 
r/z 1 1
¼ w/z þ Pq/z þ ð1  2mÞP/z  vqz ; ð40Þ
l 1m q
D. Palaniappan / Applied Mathematical Modelling 35 (2011) 5494–5499 5499

rqz 1 1
¼ ðw þ qPqqz Þ þ 2Pqz  v/z ; ð41Þ
l 1  m qz q
 
rq/ 1 1 1 Pq P 1 1
¼ wq/  2 w/ þ Pqq/  þ 2 þ 2 v//  vqq þ vq ; ð42Þ
l 1m q q q q q q
where l is the Lame’s constant.

References

[1] A.I. Lurie, Theory of Elasticity, Springer-Verlag, Berlin, Heidelberg, 2005.


[2] M.J. Boussinesq, Application des Potentiels a l’Étude de l’Équilibre et du Mouvements des Solides Élastiques, Gaythier-Villars, Paris, 1885.
[3] P.F. Papkovich, Solution generale des equations differentielles fondamentales d’elasticite experimee par trois fonctions harmoniques, C.R. Acad. Sci.
Paris 195 (1932) 513–515.
[4] H. Neuber, Ein Neuer Ansatz Zur Lösung räumblicher Probleme der Elastizitätstheorie, Z. Angew. Math. Mech. 14 (1934) 203–212.
[5] A.E.H. Love, A Treatise on the Mathematical Theory of Elasticity, fourth ed., Dover, New York, 1944.
[6] N. Phan-Thien, S. Kim, Microstructures in Elastic Media: Principles and Computational Methods, Oxford University Press, Oxford, 1994.
[7] P.M. Naghdi, C.S. Hsu, On the representation of displacements in linear elasticity in terms of three stress functions, J. Math. Mech. 10 (1961) 233–245.
[8] J. Happel, H. Brenner, Low Reynolds Number Hydrodynamics, Martinus Nijhoff, The Hague, 1983.
[9] D.D. Joseph, Helmholtz decomposition coupling rotational to irrotational flow of a viscous fluid, Proc. Natl. Acad. Sci. 103 (2006) 14272–14277.
[10] R. Muki, Asymmetric problems of the theory of elasticity for a semi-infinite solid and a thick plate, in: I.N. Sneddon, R. Hill (Eds.), Progress in Solid
Mechanics, North-Holland Publish Co., Amsterdam, 1960, pp. 401–430.
[11] J.G. Simmonds, Love’s stress function for torsionless axisymmetric deformation of elastically isotropic bodies with body forces, ASME J. Appl. Mech. 67
(2000) 628–629.
[12] T. Tran-Cong, On the solutions of Boussinesq, Love, and Reissner and Wennagel for axisymmetric elastic deformations, Q. J. Mech. Appl. Math. 50
(1997) 195–210.
[13] M.Z. Wang, X.X. Xu, A generalization of Almansi’s theorem and its application, Appl. Math. Model. 14 (1990) 275–279.
[14] J.H. Michell, The uniform torsion and flexure of incomplete tores, with application to helical springs, Proc. Lond. Math. Soc. 31 (1900) 130–146.
[15] Y.C. Lou, M.Z. Wang, Michell’s general solutions for torsionless axisymmetric problems with body forces in elasticity, ASME J. Appl. Mech. 70 (2003)
448–449.
[16] P. Burgatti, Sztpra due utili forme dell’integrale generale, dell’equazioni per l’equilibrio dei solidi elaetici isotropi, Mem. R. Acad. Sci. VIII3 (1926) 63–67.
[17] A.E. Green, W. Zerna, Theoretical Elasticity, Oxford University Press, Oxford, 1968. p. 165.

You might also like