You are on page 1of 26

INTERNATIONAL JOURNAL FOR NUMERICAL METHODS IN ENGINEERING, VOL.

41, 473—498 (1998)

AN ACCURATE TWO-NODE FINITE ELEMENT FOR


SHEAR DEFORMABLE CURVED BEAMS

Z. FRIEDMAN AND J. B. KOSMATKA*


Department of Applied Mechanics and Engineering Sciences, ºniversity of California, San Diego, ¸a Jolla, CA 92093, º.S.A.

ABSTRACT
An accurate two-node (three degrees of freedom per node) finite element is developed for curved shear
deformable beams. The element formulation is based on shape functions that satisfy the homogeneous form
of the partial differential equations of motion which renders it free of shear and membrane locking. The
element is demonstrated to converge to the results obtained from a shear deformable straight beam when the
beam becomes shallower. Numerical examples were performed to demonstrate the accuracy and efficiency
with respect to previously published formulations. ( 1998 John Wiley & Sons, Ltd.
Int. J. Numer. Meth. Engng., 41, 473—498 (1998)

KEY WORDS: shear deformable; curved beams; two node finite element

INTRODUCTION
A common practice in the finite element analysis of a curved geometry is the use of a large number
of straight elements to approximate the true curved shape. For better efficiency, special elements
for curved geometry in which the coupling between the transverse and axial motions is captured
through the element formulation were developed. These elements are important not only for
modelling 1-D curved geometry but also as a precursor for the development of a 2-D curved shell
element. Early attempts to formulate a curved finite beam element were unsuccessful and resulted
in an inaccurate element that converged more slowly than straight beam approximation to the
curved geometry.1 Early literature1, 2 felt that the poor performance was due to the improper
description of the rigid body modes. More recently, researchers3 felt that the poor behaviour was
due to both shear and membrane locking, where an overly stiff behaviour of a curved beam is due
to large axial stiffness relative to the bending stiffness. The membrane locking phenomenon
becomes less severe as the element becomes shallower. As in the straight beam situation, the
methods that have been used to alleviate these locking phenomena are the use of mixed
formulations or the use of reduced integration. Recently higher-order polynomials has been
successfully used for shape functions to alleviate the locking phenomenon but they required more
nodes per element or more degrees of freedom per node.3—9 Some of these formulations4 use
higher-order polynomials with constraints to limit the number of nodes and/or degrees of
freedom used. The most accurate element published to date is the one developed in Reference 4

*Correspondence to: J. B. Kosmatka, Department of Applied Mechanics and Engineering Sciences, University of
California, San Diego, La Jolla, CA 92093-0085, U.S.A.

CCC 0029—5981/98/030473—26$17.50 Received 24 February 1997


( 1998 John Wiley & Sons, Ltd. Revised 9 May 1997
474 Z. FRIEDMAN AND J. B. KOSMATKA

which is a three-node element. In spite of the fact that this element is free of membrane and shear
locking it still suffers from displacement inaccuracies under concentrated loads.
In this article, an accurate two-node (three degrees of freedom per node) curved finite beam
element is developed. The element formulation is based on shape functions that satisfy the
homogeneous form of the partial differential equations of motion which renders it free of shear
and membrane locking. The element is demonstrated to converge to the results obtained from
a shear deformable straight beam when the beam becomes shallower. Numerical examples
demonstrate the accuracy and efficiency with respect to previously published formulations.

THEORETICAL DEVELOPMENT
The development of a shear deformable curved beam theory is done along the same lines as the
development of the theory for a straight (Timoshenko) beam. The main differences are that for the
curved beam the development is performed in a curvilinear co-ordinate system and there is an
inherent coupling between the axial and transverse motions. This coupling is the source for the
elements efficiency but it also requires three unknown functions (transverse, rotational and axial
motions) that are coupled in the differential equations rather than two unknown functions only
(transverse and rotational motions) in Timoshenko’s theory for straight beam.
Figure 1 depicts a prismatic isotropic curved beam having a general cross-section of area A and
moment of inertia I about the area’s centroid and radius of curvature R. We define a curvilinear
coordinate system (x, y, z) such that the x-co-ordinate is coincident with the centroidal curved
axis and y and z are coincident with the principal axes of the cross-section. The centroidal axis
(x-co-ordinate) is the line about which the cross-sections are rotating during bending and
therefore represents a zero stress point at each cross-section and is known as the neutral axis. In
some texts of advanced strength of materials10 the location of neutral axis is displaced from the
centroid of the cross-section towards the centre of curvature accounting for the fact that the fibres
closer to the centre of curvature are initially shorter than those farther away from the centre of
curvature. This gives a hyperbolic axial stress distribution in the cross-section rather than linear
which is more accurate especially for short deep beams. The location of the neutral axis in the

Figure 1. Curved beam element

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 475

cross-section does not affect the development that follows which will be performed for motions in
the x—y plane only since it is assumed that no bending-torsion coupling exists due to the choice of
y and z in the principal directions of the cross-section.
For the shear-deformable curved beam the displacements are prescribed as follows:

u(x, y, z, t)"u(x, t)!yh(x, t), v(x, y, z, t)"v(x, t), w(x, y, z, t)"0 (1a—c)

where u, v and w are the axial, radial and out of plane displacements of the centroidal axis,
respectively, and h is the rotation of the cross-section about the z-axis.
The strain components of the beam are determined using the definitions for strain in a cur-
vilinear (polar) co-ordinate system and using equations (1a)—(1c). The two non-zero strain
components are:

Lu v Lh Lv u
e " ! !y , c " !h# (2a, b)
xx Lx R Lx xy Lx R

The equations of motions will be developed using Hamilton’s principle:

P
t
2
d%" (dº!d¹!d¼ ) dt"0 (3)
t
%
1

where dº, d¹ and d¼ are the variations of the strain energy, the kinetic energy, and the work of
%
external forces, respectively. The strain energy is given by:

1 L
º"
2
0 A
PP
MpNTMeN dA dx (4a)

and can be rewritten, by making use of equations (2a) and (2b), the beam constitutive relations,
and integrating over the cross-section:

G H G H
Lu v T Lu v
! !
Lx R EA 0 0 Lx R
1 L Lh Lh
º"
2
0
P Lx
0 EI 0
Lx
dx (4b)

Lv u 0 0 kGA Lv u
!h# !h#
Lx R Lx R

where E and G are Young’s and shear moduli of the beam material, respectively and k is a shear
coefficient that is dependent upon the cross-section geometry.11 Since the integration is performed
along the beams length, ¸ is the total arc-length of the curved beam. Equation (4b) can be written
in matrix notation as follows:

1 L
º"
2
0
0 P
[¹ ]T[D][¹ ] dx
0
(4c)

( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
476 Z. FRIEDMAN AND J. B. KOSMATKA

with the following definitions for the matrices:


L 1
! 0
Lx R

GH
u
L
[¹ ]" 0 0 v (4d)
0 Lx
1 L h
!1
R Lx

EA 0 0
[D]" 0 EI 0 (4e)
0 0 kGA
The kinetic energy of the beam is given by:

P P CA B A B A B D
1 L Lu 2 Lv 2 Lw 2
¹" o # # dA dx (5a)
2 Lt Lt Lt
0 A
The above can also be rewritten by integrating over the cross-section, as

GH GH
Lu T Lu
Lt oA 0 0 Lt
1 L Lv Lv
¹"
2
0
P Lt
0 oA 0
0 0 oI
Lt
dx (5b)

Lh Lh
Lt Lt
where o is the mass density of the beam’s material. Finally, the work of the external forces is given
by:

GHG H
u T f
L
¼"
% P0 v
h
q
m
dx (6)

where f, q and m are the distributed axial force, distributed radial force and distributed moment
along the length of the beam (given as a function of the curvilinear co-ordinate along the beam
length), respectively.
The three differential equations of motion and the associated boundary conditions are
obtained by substituting equations (4b), (5b), and (6) into (3) and integrating by parts:

A B
L2 kGA EA#kGA Lv kGA L2u
!EA # u# ! h#f"oA
Lx2 R2 R Lx R Lt2

A B
EA#kGA L2 EA L2v
u# kGA ! v!kGAh#q"oA (7a—c)
R Lx2 R2 Lt2

A B
kGA Lv L2 L2h
u#kGA # EI !kGA h#m"oI
R Lx Lx2 Lt2

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 477

and the two geometric and natural boundary conditions that must be satisfied at the beam ends
(x"0, ¸) are:

Geometric Natural

A B
u Lu v
F"EA !
Lx R
(7d)
A B
v Lv u
Q"kGA !h#
Lx R
h Lh
M"EI
Lx

Here F, Q and M are defined as the effective cross-sectional axial, shear force and bending
moment, respectively.
For the finite element formulation we will use special shape functions for u, v and h which have
the property that they exactly satisfy the homogeneous form of the static equations of equilibrium
of the curved beam:

A B
L2 kGA EA#kGA Lv kGA
!EA # u# ! h"0
Lx2 R2 R Lx R

A B
EA#kGA L2 EA
u# kGA ! v!kGAh"0 (8a—c)
R Lx2 R2

A B
kGA Lv L2
u#kGA # EI !kGA h"0
R Lx Lx2
A methodology analogous to the one used in Reference 12 (for obtaining the solution for the
straight beam equations) would be assuming a polynomial of a certain order and determining the
values of the coefficients by substitution into equations (8). This method will not be successful for
a curved beam because the solutions are not necessarily all polynomials. A more systematic
approach by which all the solutions to the differential equations will be found is assuming an
exponential function as follows:
u"aerx, v"berx, h"cerx (9a—c)
Substitution of equations (9) into equations (8) and dividing through by the exponential
function gives three homogeneous algebraic equations for the three coefficients a, b and c in
equations (9) and the parameter r is undetermined:

A B
kGA EA#kGA kGA
!EAr2# a# rb! c"0
R2 R R

A B
EA#kGA EA
a# kGAr2! b!kGAc"0 (10a—c)
R R2
kGA
a#kGArb#(EIr2!kGA)c"0
R

( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
478 Z. FRIEDMAN AND J. B. KOSMATKA

In order for a non-trivial solution to exist for a, b and c the matrix of the coefficients must be
singular, namely its determinant must be equal to zero. This gives an equation for r. After some
simplification the equation for r becomes:

2 1
r6# r4# r2"0. (11a)
R2 R4

Equation (11a) can be rewritten as:

A B
1 2
r2 r2# "0 (11b)
R2

Note that the equation for r is independent of the elastic properties and of the cross-sectional
area of the beam. That basically means that the characteristics of the solution being trigonometric
functions of r (see equations (12)) are independent of these properties. The values of the
coefficients of the trigonometric functions do depend on these properties (see equations (13)
and (14)).
The first and most obvious solution for r is zero and that means that a constant value is
a solution to the differential equations. But since r is squared that means a multiplicity and
requires that the linear polynomial x must be also a solution. The second family of solutions is
based on the solution for r according to the expression inside the brackets in equations (11b) and
is given by r"$(1/R) i where i"J!1. This indicates that the general solutions to equations
(8a)—(8c) consists of trigonometric functions (sine and cosine) of the argument x/R. Moreover, The
fact that the expression in the brackets of equation (11b) is squared is an indication of a multipli-
city and requires that x times the trigonometric functions will also be solutions to equations
(8a)—(8c). Therefore, the functions u, v, and h must be of the following form:

AB AB AB AB
x x x x
v"a #a x#a cos #a x cos #a sin #a x sin
1 2 3 R 4 R 5 R 6 R

AB AB AB AB
x x x x
h"b #b x#b cos #b x cos #b sin #b x sin (12a—c)
1 2 3 R 4 R 5 R 6 R

AB AB AB AB
x x x x
u"c #c x#c cos #c x cos #c sin #c x sin
1 2 3 R 4 R 5 R 6 R

Equations (12a)—(12c) can be expressed in matrix form as follows:

GH G H
u [N ] 0 0 McN
1
v " 0 [N ] 0 MaN (12d)
1
h 0 0 [N ] MbN
1

with the following definitions for the matrices:

C D
x x x x
[N ]" 1 x cos x cos sin x sin (12e)
1 R R R R

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 479

MaNT"Ma a a a a a N
1 2 3 4 5 6
MbNT"Mb b b b b b N (12f—h)
1 2 3 4 5 6
McNT"Mc c c c c c N
1 2 3 4 5 6
The constants a , b and c where i"1, . . . , 6 are not all independent. By substituting equations
i i i
(12a)—(12c) into equations (8a)—(8c) and requiring them to vanish identically we can obtain 12
constraints which will leave exactly six independent coefficients (i.e. there are six boundary
conditions). It was found that the constraints can be expressed in their simplest form if the six
coefficients chosen as independent are a , a , a , c , c , c . These coefficients will later on have to
1 3 5 1 3 5
be expressed in terms of the boundary conditions at the beam’s ends.
The 12 constraints that allow the rest of the coefficients to be expressed as functions of the
independent coefficients are:
k k
a "0, a " 2 (a #c ), a " 2 (!a #c )
2 4 k R 5 3 6 k R 3 5
1 1
c a 2
b " 1, b " 1 , b " (a #c )
1 R 2 R2 3 k R 5 3
1 (13a—l)
2
b "0, b " (!a #c ), b "0
4 5 k R 3 5 6
1
a k k
c " 1 , c "! 2 (!a #c ), c " 2 (a #c )
2 R 4 k R 3 5 6 k R 5 3
1 1
Note that three of the coefficients above must be zero in order for this set of displacements to
satisfy the homogeneous set of differential equations, equations (8). This fact is of no special
significance since it just means that certain terms (e.g. the linear term for the transverse motion)
can never be part of the solution.
The relationships in equations (13) can be written in matrix notation as follows:

G H
McN
MaN "[¹ ]MdN (13m)
MbN
with the following definitions:
MdNT"Ma a a c c c N (13n)
1 3 5 1 3 5
0 1/R 0 0 0 0 1 0 0 0 0 0 0 1/R2 0 0 0 0
0 0 0 !k1 0 0 0 0 1 0 0 !k1 0 0 0 0 !kª 0
0 0 0 0 0 kM 0 0 0 k1 1 0 0 0 kª 0 0 0
[¹]T" (13p)
1 0 0 0 0 0 0 0 0 0 0 0 1/R 0 0 0 0 0
0 0 1 0 0 k1 0 0 0 k1 0 0 0 0 kª 0 0 0
0 0 0 k1 1 0 0 0 0 0 0 k1 0 0 0 0 k1 0

( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
480 Z. FRIEDMAN AND J. B. KOSMATKA

where the constants in the above matrices are defined by

A BA B
kGA EI ¸2
k "1# 1!
1 EA kGA¸2 R2

A BA B
kGA EI ¸2
k "1# 1#
2 EA kGA¸2 R2

k
k1 " 2
k R
1
2
kª " (14a—d)
k R
1
Note that in the expressions for k and k , the terms inside the first and the second parentheses
1 2
are related to different ratios between the axial, bending and shear stiffness properties of the
beam.
The stiffness matrix defined in terms of the generalized coordinates MdN is obtained by
substituting equations (12d), (13m), (4d) and (4c) into Hamilton’s principle—equations (3).

AP B
#
[Ka]"[¹ ]T [B]T [D][B]R dh [¹] (14e)
0
where the matrix [¹ ] is defined in equation (13p) and the matrix [B] is defined by

L 1
[N ] ! [N ] 0
Lx 1 R 1
L
[B]" 0 0 [N ] (14f )
Lx 1
1 L
[N ] [N ] ![N ]
R 1 Lx 1 1

The integration is performed along the arc-length and # is the total included angle.
The integration is performed explicitly because the integrands contain both trigonometric and
polynomial functions. Since the matrix is symmetric, only the upper half is required to be
calculated. The elements of the matrix are as follows:

k2 #
Ka " 1 C
11 2 be

AB
#
Ka "!2k sin C
12 1 2 be

Ka "Ka "Ka "0


13 14 15

AB
#
Ka "2k sin C
16 1 2 be

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 481

Ka "(C #C #C )##(C !C #C ) sin(#)


22 ax sh be ax sh be
Ka "Ka "Ka "0
23 24 25
Ka "!(C #C #C )#!(C !C #C ) sin(#)
26 ax sh be ax sh be
Ka "(C #C #C )#!(C !C #C ) sin(#)
33 ax sh be ax sh be
Ka "0
34
Ka "(C #C #C )#!(C !C #C ) sin(#)
35 ax sh be ax sh be
Ka "0
36
Ka "Ka "Ka "0
44 45 46
Ka "(C #C #C )#!(C !C #C ) sin(#)
55 ax sh be ax sh be
Ka "0
56
Ka "(C #C #C )##(C !C #C ) sin(#) (15)
66 ax sh be ax sh be
where again # is the total included angle of the curved beam and the constants C , C and
ax sh
C are defined as follows:
be
EA(k !k )2
C " 1 2
ax 2k2 R
1
kGA(k #k !2)2
C " 1 2
sh 2k2 R
1
2EI
C " (16a—c)
be k2 R3
1
The constant k and k are defined in equations (14a) and (14b)
1 2
The generalized force vector in the generalized co-ordinates is obtained in the same manner.
Using equations (12d), (13m) and (6) to substitute into Hamilton’s principle—equations (3)—we
can express the 6]1 force vector in the generalized co-ordinates MdNas follows:

A BG H
[N ] 0 0 T
1 f
#
MFaN"[¹ ]T
P0 0 [N ]
1
0 R dh
m
q (17a)

0 0 [N ]
1
where f, q and m are the distributed axial force, distributed radial force and distributed moment
along the length of the beam (given as a function of the curvilinear co-ordinate along the beam
length). Explicit integration yields the elements of the generalized force as follows:
Fa "R#q
1
# #
k # cos #2(k !k ) sin
2 2 1 2 2
Fa "R q
2 k
1
( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
482 Z. FRIEDMAN AND J. B. KOSMATKA

A B
# # #
k # cos !2 sin 4 sin
2 2 2 2
Fa "!R f# m
3 k k
1 1
Fa "R# f##m
4
# # #
!k # cos #2(k #k ) sin 4 sin
2 2 1 2 2 2
Fa "R f# m
5 k k
1 1

A B
# #
k # cos !2 sin
2 2 2
Fa "!R f (17b—g)
6 k
1
where all the constants used were previously defined.
The development of the mass matrix requires a slightly different approach. The direct use of the
chosen independent coefficients a , a , a , c , c , c as was done for the stiffness matrix will result
1 3 5 1 3 5
in very long and complicated expressions for the mass coefficients. According to equation (5b) the
kinetic energy is a summation of the contributions of the three motions (transverse, axial and
rotations). Therefore, the three mass matrices will be developed in their own generalized
co-ordinates and than transformed to the six independent coefficients mentioned above.
Since the general expressions for all three displacements in equations (12) are identical (with
different coefficients), the mass matrix [Mg] in the generalized co-ordinates of each one of them
will be identical if we ignore the multiplying factor of oA or oI. This matrix can be obtained by
using equation (5b) for the kinetic energy and performing the integrations. The components of
this matrix are:

AB
#
Mg "R#, Mg "2R sin , Mg "Mg "Mg "0
11 13 2 12 14 15

A AB A BB
# #
Mg "R2 2 sin !# cos
16 2 2
R3#3
Mg " , Mg "Mg "0
22 12 23 26

A AB AB A BB
# # #
R3 4# cos !8 sin ##2 sin
2 2 2
Mg "
24 2

A AB A BB
# #
Mg "R2 2 sin !# cos
25 2 2
R(##sin(#)) R2(sin(#)!# cos(#))
Mg " , Mg "Mg "0, Mg "
33 2 34 35 36 4
R3(#3#6# cos(#)!6 sin(#)#3#2 sin(#))
Mg "
44 24
R2(sin(#)!# cos(#))
Mg " , Mg "0
45 4 46

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 483

R(#!sin(#))
Mg " , Mg "0
55 2 56

R3(#3!6# cos(#)#6 sin(#)!3#2 sin(#))


Mg " (18)
66 24

Note that the matrix [Mg] is symmetric therefore only its upper half is given above. # and R are
as before the total included angle and the radius of curvature of the beam, respectively.
The specific mass matrices for each one of three displacements (transverse, axial and rotational)
can be obtained by transformation from their own generalized co-ordinates to the six indepen-
dent variables a , a , a , c , c , c that were previously chosen. Therefore,
1 3 5 1 3 5
For the transverse motion: [M ]"oA[aN ]T[Mg][aN ]
V
For the axial motion: [M ]"oA[bM ]T[Mg][bM ]
U
For the rotational motion: [M ]"oI[cN ]T[Mg][cN ] (19a—c)
h
The matrices in equations (19) are determined by the relationships between the coefficients
defined in equations (13). They are obtained as follows:

1 0 0 0 0 0
0 0 0 0 0 0
0 1 0 0 0 0
[aN ]" (20)
0 0 k /(k R) 0 k /(k R) 0
2 1 2 1
0 0 1 0 0 0
0 !k /(k R) 0 0 0 k /(k R)
2 1 2 1

0 0 0 1/R 0 0
1/R2 0 0 0 0 0
0 0 2/(k R) 0 2/(k R) 0
[b1 ]" 1 1 (21)
0 0 0 0 0 0
0 !2/(k R) 0 0 0 2/(k R)
1 1
0 0 0 0 0 0

0 0 0 1 0 0
1/R 0 0 0 0 0
0 0 0 0 1 0
[cN ]" (22)
0 k /(k R) 0 0 0 !k /(k R)
2 1 2 1
0 0 0 0 0 1
0 0 k /(k R) 0 k /(k R) 0
2 1 2 1
( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
484 Z. FRIEDMAN AND J. B. KOSMATKA

The mass matrix [Ma] in the generalized co-ordinates a , a , a , c , c , c will therefore be given
1 3 5 1 3 5
by
[Ma]"[M ]#[M ]#[M ] (23)
U V h
The transformation to the physical co-ordinates is performed using the usual transformation of
vectors and matrices between co-ordinate systems. If [K], [M] and MFN are the stiffness matrix,
mass matrix and the force vector in the physical co-ordinates, then the transformation from the
generalized co-ordinates to the physical co-ordinates are given by:
[K]"[N]T[Ka][N], [M]"[N]T[Ma][N], MFN"[N]TMFaN (24)
where the transformation matrix [N] is the matrix that is used to express the generalized
co-ordinates in term of the physical co-ordinates. We define the following vectors:
MaN"Ma a a c c c NT
1 3 5 1 3 5
MwN"Mv h u v h u NT (25a, b)
1 1 1 2 2 2
where MaN is the vector of generalized co-ordinates and its components are the independent
coefficients in the expressions for displacements according to equations (12) and isMwN the vector
of physical co-ordinates and its components are the physical motions at the beam’s ends (see
Figure 1). Then [N] is defined by the following relationship:
MaN"[N]MwN, MwN"[NM ]MaN, [N]"[NM ]~1 (26a—c)
The actual components of matrix [NM ] can be obtained by substitution of the values of the two
ends of the beam in equations (12). This matrix has to be inverted in order to obtain [N]. The
explicit form of matrix [NM ] is

cos #!k
1 sin ##k 0 k k
3 4 4 3
#/R !k k 1/R k k
3 6 6 5
# k k 1 cos ##k sin #!k
[NM ]" 4 3 3 4 (27)
1 cos #!k !sin #!k 0 0 k
3 4 3
!#/R k k 1/R k !k
5 6 6 5
!# !k k 1 cos ##k !sin ##k
4 3 3 4
where
k k
k " 2 # sin # k " 2 # cos #
3 k 4 k
1 1
2 2
k " sin # k " cos # (28a—d)
5 k R 6 k R
1 1

NUMERICAL STUDIES
The validity and efficiency of the element formulated in the previous section are verified through
numerical tests. Both static and dynamic numerical tests were performed. For the static analysis,
comparison was made with existing curved elements.

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 485

Static analysis
The current element must converge to Timoshenko’s straight beam results as the element
becomes shallower, since its formulation includes shear deformation. A series of tests were
performed to prove this convergence. A measure for element shallowness, named ‘sag ratio’, was
defined as the ratio of half-span of the element to the sag at the middle of the element. See
Figure 2(a) for details. The convergence was studied by varying the sag ratio from the value of
unity (which is equivalent to a very deep curved beam with included angle of 180°) to the value of
10 000 (very close to a straight beam). Curved beams with different sag ratios are illustrated
in Figure 2(a).
A cantilever beam with five different loading conditions was used for this numerical study. The
loading conditions applied were: tip radial force, tip axial force, tip moment, distributed radial
force and distributed moment. See Figures 2(b) and 2(c) for a depiction of the five loading
conditions. The three tip displacements (i.e. radial, rotational and axial) for the five loading
conditions were calculated using only one element and compared to the theoretical Timoshenko’s
straight beam solutions. In each case the results of the finite element analysis were normalized to
Timoshenko’s straight beam analytical results in order to obtain a non-dimensional relative
displacements which were expected to converge to the value of unity. Obviously, this could not be
done for Timoshenko’s analytical results that are theoretically zero. In these cases, the absolute
displacement were studied and were expected to converge to zero.

Figure 2. Curved beam element: (a) different sag ratios, (b) tip loading, (c) distributed loading

( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
486 Z. FRIEDMAN AND J. B. KOSMATKA

The following geometrical and material parameters were used for this study:
Young’s modulus (E)"30]106
Shear modulus (G)"10]106
Shear correction factor (k)"0·85
Cross-sectional area (A)"25
Cross-sectional moment of inertia (I)"50
Beam’s span (¸)"100
The cross-sectional properties approximately represent a square of 5]5. Therefore, the above
properties represent a relatively slender beam (length/depth"20) which can detect a locking
phenomenon. The radius of curvature and the total included angle varied with the sag ratio’s
values. For the sag ratio varying from 1 to 10 000, the radius of curvature varied from 50 to
250 000 and the total included angle varied from 180° to 0·023°. The numerical results are
depicted in Figure 3. The tip’s relative radial, rotational and axial displacements converged to the
value of unity and the absolute displacements converged to zero as expected. Notice that
convergence occurred with one element only which is an indication to the element’s accuracy and
efficiency.
The next study was performed to prove that the element is free of shear and membrane locking.
In this study, a very shallow one beam element with sag ratio of 1000 was used (radius of
curvature was 25 000, total included angle was 0·23° and the span was of 100). All the geometrical
and material properties in the previous study were used here except for the cross-sectional
properties that were changed to achieve different stiffness ratios as described below. The shear
locking phenomenon is manifested in an overly stiff response (locking) to loading and it occurs as
the beam becomes thinner such that the shear stiffness becomes much larger than the bending
stiffness. The same kind of behaviour is characterized by membrane locking which occurs as
the axial stiffness becomes much larger than the bending stiffness. Both the shear stiffness and
the axial stiffness are in direct proportion to the cross-sectional area while the bending stiffness
is in direct proportion to the cross-sectional moment of inertia. Therefore, a non-dimensional
‘slenderness ratio’ to capture both the bending to shear stiffness ratio and bending to axial
stiffness ratio can be defined as follows:

A B
I R 2
Slenderness Ratio"a" " ' (29)
¸2A ¸
where R is the radius of gyration of the cross-section. In order to test the possibility of shear and
'
membrane locking the value of the slenderness ratio was varied from 0·1 (short deep beam) to
0·000001 (long slender beam). This was achieved by varying the value of the cross-sectional area
from 0·05 to 5000 while retaining the cross-sectional moment of inertia and the length constant.
As the magnitude of the shear and axial stiffness is increased the tip radial displacement for this
almost straight beam is expected to converge to the Euler—Bernoulli’s analytical solution. For the
numerical analysis a cantilever beam with one finite element was used and two types of shear
loading were applied: tip concentrated shear force and distributed shear load. The numerical
results were normalized to the Euler—Bernoulli’s analytical solution.

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 487

Figure 3. Convergence of current curved element to Timoshenko beam’s tip displacements

The numerical results are depicted in Figure 4. It is readily observed that the tip radial
displacements for both loading conditions converged to the correct values and no locking was
observed.
Since the tendency to exhibit membrane locking is stronger for a deep curved beam than for
a shallow beam, the study above was repeated for a deep curved beam with a sag ratio of 1·0

( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
488 Z. FRIEDMAN AND J. B. KOSMATKA

Figure 4. Convergence of current curved element to Euler—Bernoulli’s beam’s tip radial displacement

(included angle of 180°). The numerical results for the tip radial displacement with one finite
element were exact (compared to the analytical solution based on the curved beam theory) for
both loading conditions and for any value of slenderness ratio as defined in equation (29). Thus,
the element is free of shear locking, free of membrane locking and very accurate (one element gives
exact results for concentrated and distributed load).
The next numerical study was aimed at demonstrating the element efficiency by comparing the
results of using the current curved element formulation to the results obtained by using a varying
number of straight elements to describe the curved geometry. This method is commonly used to
analyse curved geometry. The straight element used for comparison is a two-node beam element
based on Timoshenko’s beam theory13,14 and uses a correction factor called ‘residual stress
factor’.13 For this study an extremely curved beam with an included angle of 350° was used. See
Figure 5 for details. This beam is basically an almost full circle open ring. The cross-sectional
geometry and material properties were identical to the those used in the first numerical study. The
numerical results for the tip radial, axial and rotational displacements were calculated for the
cantilevered ring under four loading conditions; tip radial force, tip axial force, tip moment and
distributed radial force. See Figure 5 for a depiction of the loading conditions. A varying number
of elements (from 1 to 256) was used and the results were normalized to the exact analytical
solutions (calculated from the original differential equations). Figure 6 depicts the convergence

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 489

Figure 5. 350° cantileverd ring for numerical study: (a) tip loading, (b) distributed loading

rate to the exact solutions for the four loading conditions. Notice that for all cases the current
curved beam element gave exact results with any number of elements (including one element),
even for the distributed loading condition. This behaviour is similar to the Timoshenko based
straight beam element developed in a previous paper,12 which is expected because both elements
were developed using the same approach. The efficiency of modelling curved geometry with
straight elements (a common practice) varied from moderately inefficient to very inefficient. For
the tip radial force loading condition, 16 elements were required to reduce the error in the tip
radial and rotational displacement to under 5 per cent and a remarkable large number of
elements (128) were required to reduce the tip axial displacement to the same error level under the
same loading conditions. The same kind of behaviour was exhibited under the other loading
conditions but not necessarily for the same displacements types.
For the same curved beam (350° ring) the internal radial and axial forces at the clamped end
(the root) were calculated using equations (7d) and compared to the results obtained using the
straight element described above. Figure 7 depicts the results for loading by tip radial and axial
forces. The results were normalized to the exact values calculated from static equilibrium. As can
be observed the current curved element gave exact results for any number of elements (including
one element) while 32 straight elements were required to achieve better than 5 per cent accuracy
for the internal radial (shear) force and even 256 element were not sufficient to achieve the same
accuracy for the internal axial force.

( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
490 Z. FRIEDMAN AND J. B. KOSMATKA

Figure 6. Normalized tip displacements of a 350° cantilevered ring

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 491

Figure 7. Normalized internal forces at the root of 350° cantilevered ring

The performance of the current element was next compared to existing curved elements in
literature. First the current element was compared to four previously developed elements;3,8 two
from Reference 3 and two from Reference 8. The elements in Reference 3 are two-node elements
and those in Reference 8 are three-node elements. The results reported in the above references
start with a minimum of nine degrees of freedom namely, two elements. Two curved beam
structures were studied; a shallow thin arch and a deep thick arch. See Figures 8 and 9 for details.
The shallow and thin arch had an included angle of 15° and a ratio of the total length to
cross-section height of 100. The deep and thick arch had an included angle of 90° and a ratio of the
total length to cross-section height of 5. The beams were clamped at one end and were guided in the
radial direction at the other end. A force load was applied in the radial direction at the guided end of
the beam and the results for radial displacement at that end were compared. The results for the all
five elements are depicted for the shallow and deep beams in Figures 8 and 9, respectively. All the
displacements were normalized to the exact results obtained from the solution of the differential
equations. As shown in these figures the current element, being exact for any number of elements
(including one element), is superior to all existing elements for both types of beams.
The next study was a performance comparison of the current element with the one developed in
Reference 4 which is the most advanced previously developed curved element. It is a three node
element while the current one is a two-node element. Two beam structures were compared.

( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
492 Z. FRIEDMAN AND J. B. KOSMATKA

Figure 8. Shallow thin arch—comparison with previously published curved elements

The first beam studied was a cantilevered quarter ring loaded with a radial force applied at the
free end. See Figure 10 for details. The radial and axial tip displacements at the free end were
obtained and compared to those obtained in Reference 4.
The second structure studied was a pinched ring in which a full ring was deformed by a two
equal and opposing forces. For modelling purpose, only a quarter of the ring was analysed with
the appropriate boundary conditions (radial guidance at both ends). See Figure 11 for details. The
radial displacements at both guided ends were obtained and compared to those obtained in
Reference 4.

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 493

Figure 9. Deep thick arch—comparison with previously published curved elements

In both structures, one element was used for both type of finite elements and the ratio of the
cross-section depth to the radius of curvature was varied from 0·001 to 0·5. Figures 10 and 11
depict the results of all displacements and are normalized to the exact solution obtained from the
differential equations. As shown in the figures the current element gives exact results for all cases
and therefore superior to the element in Reference 4 which gives accurate (though not exact)
results only for thin beams.

Dynamic analysis
Two studies were aimed at showing the convergence of a deep curved beam with included angle
of 180° to the exact results as the number of elements used increased. Beams with two types of

( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
494 Z. FRIEDMAN AND J. B. KOSMATKA

Figure 10. Cantilevered 1/4 ring—comparison with previously published curved elements

boundary conditions were studied; fixed and simply supported on both ends. For both beams
comparison was made with the convergence of a straight beam element14 to the exact results. The
exact results were estimated by using a very large number (8192) of straight elements.14 The
number of elements used varied from 8 to 32. Figures 12 and 13 show the results for the fixed and
the simply supported beams, respectively. As shown in the above figures the current element
converges faster to the exact solutions than the straight element for all three natural frequency in
both beams.

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 495

Figure 11. Pinched ring—comparison with previously published curved elements

CONCLUSIONS
The methodology used in Reference 12 for a shear deformable straight element (Timoshenko) was
extended in this article to a curved beam element. Here the benefits are even larger because of the
element’s capability to model curved geometry exactly and still obtain exact results in static
analysis and better results in dynamic analysis. This capability is due to the inherent coupling
between axial and bending behaviour in the curved beam theory, but at the same time it makes
the formulation more complicated. The end result though, is a simple element (two-node/three
dof per node), free of shear and membrane locking that gives better results than previously

( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
496 Z. FRIEDMAN AND J. B. KOSMATKA

Figure 12. 180° curved beam fixed at both ends—first three vibration modes

published curved elements. The correctness of the element was proven by convergence to straight
beam’s results and its efficiency in modelling curved geometry was demonstrated by comparison
with the common practice of using many straight elements to model a curved geometry. The
methodology and results presented here can help in understanding the more complicated
behaviour of the 2-D curved shell element.

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.
SHEAR DEFORMABLE CURVED BEAMS 497

Figure 13. 180° curved beam simply supported at both ends—first three vibration modes

REFERENCES
1. J. E. Walz, R. E. Fulton, N. J. Cyrus and R. T. Eppink, ‘Accuracy of finite element approximation to structural
problems’, NASA ¹echnical Note ¹N D-5728, 1970.
2. G. Canton and R. W. Clough, ‘A curved, cylindrical-shell, finite element’, AIAA J., 6, 1057—1062 (1968).
3. G. Prathap and C. Ramesh Babu, ‘An isoparametric quadratic thick curved beam element’, Int. J. Numer. Meth.
Engng., 23, 1583—1600 (1986).
4. P.-G. Lee and H.-C. Sin, ‘Locking-free curved beam element based on curvature’, Int. J. Numer. Meth. Engng., 37,
989—1007 (1994).

( 1998 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng., 41, 473—498 (1998)
498 Z. FRIEDMAN AND J. B. KOSMATKA

5. S.-Y. Yang and H.-C. Sin, ‘Curvature-based beam elements for the analysis of Timoshenko and shear-deformable
curved beam’, J. Sound »ib., 187, 569—584 (1995).
6. C. Ramesh Babu and G. Prathap, ‘A linear thick curved beam element’` , Int. J. Numer. Meth. Engng., 23, 1313—1328
(1986).
7. G. Prathap, ‘The curved beam/deep arch/finite ring element revisited’, Int. J. Numer. Meth. Engng., 21, 389—407 (1985).
8. J. K. Choi and J. K. Lim, ‘Simple curved shear beam element’, Comm. Numer. Meth. Engng., 9, 659—669 (1993).
9. F. Kikuchi and K. Tanizawa, ‘Accuracy and locking-free property of the beam element approximation for arch
problems’, Comput. Struct., 19(1-2), 103—110 (1984).
10. R. G. Budynas, Advanced Strength and Applied Stress Analysis, McGraw-Hill, New York, 1977, p. 128.
11. G. R. Cowper, ‘The shear coefficient in Timoshenko’s beam theory’, ASME J. Appl. Mech., 33, 335—340 (1966).
12. Z. Friedman and J. B. Kosmatka, ‘An Improved Two-Node Timoshenko Beam Finite Element’, Comput. Struct.,
47(3), 473—481 (1993).
13. R. H. MacNeal, ‘A simple quadrilateral shell element’, Comput. Struct., 8, 175—183 (1976).
14. MacNeal-Schwendler Corporation, MSC/NASTRAN, Handbook for ¸inear Analysis, The MacNeal-Schwendler, 815
Colorado Blvd., Los Angeles, CA 90041, U.S.A., 1985.

Int. J. Numer. Meth. Engng., 41, 473—498 (1998) ( 1998 John Wiley & Sons, Ltd.

You might also like