You are on page 1of 9

Comput. Methods Appl. Mech. Engrg.

200 (2011) 233–241

Contents lists available at ScienceDirect

Comput. Methods Appl. Mech. Engrg.


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

Cylindrical element: Isogeometric model of continuum rod


Jia Lu a,⇑, Xianlian Zhou b
a
Department of Mechanical and Industrial Engineering, Center for Computer Aided Design, The University of Iowa, Iowa City, IA 52242-1527, USA
b
CFD Research Corporation, 215 Wynn Drive, Huntsville, AL 35805, USA

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

Article history: A geometric mapping between a canonical cylindrical domain and a physical rod-like body is established
Received 9 April 2010 and is utilized to analyze rod-like structures undergoing large deformations. A rod is described as a full
Received in revised form 12 August 2010 three-dimensional continuum, yet parameterized only by surface control points. This description pre-
Accepted 13 August 2010
serves the geometric smoothness, and allows for the consideration of lateral deformation in analysis.
Available online 20 August 2010
Numerical examples are presented to demonstrate the method.
 2010 Elsevier B.V. All rights reserved.
Keywords:
Isogeometric analysis
Nonlinear rod
NURBS
Harmonic mapping
Harmonic coordinates

1. Introduction in geometry. Here, a rod body is described by a combination of


Computer Aided Design (CAD) and Computer Graphics (CG) geom-
A rod is a slender three-dimensional body for which the cross etries. Dimensional reduction is achieved at the geometric level, by
sectional dimensions are much smaller than the longitudinal utilizing a surface description involving only surface control points.
dimension. In classical mechanics, a rod is described as a deform- In a recent publication, the first author developed a method for
able curve with some small cross sectional area around it. The parameterizing a smooth surface using boundary control points
Kirchhoff–Love theory treated a rod as an inextensible curve with [6]. In that description, the NURBS representation of a boundary
rigid cross sections that remain perpendicular to the centerline curve is extended into the interior, resulting in a harmonic map-
during deformation [1]. Antman [2] generalized the Kirchhoff– ping between the canonical unit disk and a smooth 2D domain.
Love theory to allow for longitudinal extensions and transverse Combining the cross section mapping with a NURBS centerline
shear, but the cross sections are assumed to remain rigid. Based leads to a tensor-product rod, whereby a point in the rod is repre-
on Antman’s theory, many researchers developed numerical for- sented as a linear combination of the surface control points. This
mulations for analyzing rod structures, see e.g. [3,4], and [5, representation is utilized in the isogeometric framework for anal-
Chapter 17]. ysis. The method has the following attributes:
The afore-mentioned theories share a common feature that the
rod cross sections are not allowed to deform. This assumption re- 1. Geometry and analysis are placed on a uniform basis.
duces a rod body to a one-dimensional continuum and thus signif- 2. Only surface control points are needed.
icantly simplifies the equilibrium theory. However, it precludes the 3. Smooth geometries are preserved. Some common geometries
consideration of lateral deformations. In practical applications, in can be described exactly.
particular in the analysis of soft biological structures, one often
encounters slender bodies undergoing lateral deformations or sub- The Cosserat theory of rod [7–11] allows for cross sectional
ject to lateral contact. Geometrically, these systems are best mod- deformations. Rubin developed a numerical formulation for Coss-
eled as rods, and yet kinetically, in the current paradigm they will erat rod [12]. A novel 3D finite element based on the concept of
have to be described by a full three-dimensional theory. Cosserat point was also proposed [13]. The challenge of Cosserat
This paper proposes a new numerical formulation for rod-like approaches lies in the description of geometry, namely, how to
structures. In this approach, a rod is described as a three-dimen- parameterize a smooth rod using directors. The implementations
sional continuum as it eventually is. The point of departure lies in [12,13] relate the directors to element nodal vectors and essen-
tially utilize the finite element geometry, which is not effective for
⇑ Corresponding author. Tel.: +1 319 3356405; fax: +1 319 3355669. describing smooth geometries. In this regard, this paper may
E-mail address: jia-lu@uiowa.edu (J. Lu). suggest a different geometric basis for Cosserat rod formulations.

0045-7825/$ - see front matter  2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.cma.2010.08.007
234 J. Lu, X. Zhou / Comput. Methods Appl. Mech. Engrg. 200 (2011) 233–241

The article begins with a review of NURBS and harmonic coor- X


n
x¼ T i ðnÞPi : ð6Þ
dinates, followed by the introduction of the tensor-product cylin-
i¼1
ders. Interested readers are referred to [14] for a comprehensive
coverage on NURBS. A discussion on harmonic coordinates in a cir- Eq. (6) defines a harmonic mapping F: D ´ P that maps the unit disk
cular domain is contained in [6]. The proposed cylindrical element D to the domain P. Notably, the mapping is parameterized by
is presented in Section 3. The element equations are established in boundary control points only.
the isoparametric setting, in the same manner as the isogeometric The canonical configuration introduced in [6] is outlined below.
analysis [15–21]. Numerical examples are presented in Section 4 to The unit circle is divided into n equal arcs and each arc is repre-
demonstrate the features of the method. The work aims at soft sented by a 2nd degree Bezier curve. To ensure C1 continuity, the
body analysis; the application to biological systems is demon- joints between two adjacent arcs are eliminated. This yields the
strated using an example of skeletal muscle modeling. following boundary basis functions:
8  
>
> 1 tt i2
N2;2 ti1 t i2 6 t < t i1 ;
2. Geometric basis >
> 2 t
>
>  
i2
    
>
<1 N tt i1
2 0;2 t t þ 2N1;2 ttt i1
t i1
þ N 2;2 ttt i1
t i1
t i1 6 t < t i ;
2.1. Harmonic mapping on unit disk Ui ¼ 
i i1

i i
>
>
>
> 1
N0;2 t ttti t i 6 t < tiþ1 ;
>
> 2 iþ1 i
Reference [6] introduced a geometric mapping that maps the >
:
0 otherwise;
unit circular disk to a simply-connected smooth 2D surface. Map-
pings as such can be established in many ways; in [6] a particular ð7Þ
emphasis was given to boundary-based descriptions in order to where N0,1, N1,2, N2,2 are rational Bezier functions recorded in Eq.
avoid interior control points. The starting point is to parameterize (2) in reference [6]. This form is referred to as the reduced Bezier
the unit circle S: jgj = 1 by a rational Bezier curve: circle; an n-arc circle contains exactly n control points. The para-
X
n metric domain of the circle is taken to be [0, n]; the ith arc corre-
g¼ Ui ðtÞQ i ; ð1Þ sponds to the knot interval [ti, tj+1] = [i  1, i]. Some representative
i¼1 examples are shown in Fig. 1. Fig. 2 depicts a basis function in
here Qi are the control points and Ui are the rational Bezier bases. It the 8-point disk.
is known that rational Bezier functions can exactly describe conic
sections, and some admissible parameters for the unit circle are dis- Remark 2.1. Evaluating the basis function Ti requires numerical
cussed in [22], among others. The Bezier form (1) defines a mapping integration over the unit disk, for which a quadrature rule was
c: [a, b] ´ S which is assumed to be one-to-one. One can extend this introduced in [6]. The computation, although costly, is one-time
mapping into the interior to express a point n in the unit disk D as a only because the basis functions are defined in the canonical
linear combination of Qi. In [6], this was achieved using a harmonic domain. In implementation, the function values and derivatives at
mapping. Consider the Laplace equation in the unit circular domain. quadrature points are computed a priori outside element
If one regards a component of g as the prescribed Dirichlet bound- computations.
ary data, then, the corresponding component of the coordinate
function n is the solution of the Laplace equation. Given that such
a solution is represented by Poisson’s integral [23, Chapter 4], we Remark 2.2. The mapping (6) can generate concave cross sections
can conclude: as the one shown in Fig. 3. However, in such cases the mapping
Z may not be one-to-one, and caution must be exercised when
n¼ Hðn; gÞgdsg ; ð2Þ applying such geometries in analysis.
g2C
1jnj2
where Hðn; gÞ ¼ 21p jgnj2
is the Poisson kernel. Upon using (1), we 2.2. NURBS curve
obtain:
X
n
NURBS is widely utilized to represent curves and surfaces. A
n¼ T i ðnÞQ i ; ð3Þ NURBS curve has the form:
i¼1
Pm
where Si;p ðuÞwi Pi
Z xðuÞ ¼ Pi¼1
m ; ð8Þ
i¼1 Si;p ðuÞwi
1
T i ðnÞ :¼ Hðn; gÞUi  c dsg ð4Þ Pm S ðuÞwi
g2C or simply xðuÞ ¼ i¼1 Ri ðuÞPi with Ri ¼ Pmi;p . In (8), Pi are the
S ðuÞwi
j¼1 j;p
are the basis functions. By construction, Ti(n) are solutions of the La- control points, wi are the weights associated with the control points,
place equation and therefore are called harmonic basis, or harmonic and Si, p are the B-spline basis functions of degree p. The B-spline ba-
coordinates. On the boundary, Ti = Ui c1 (Kronecker delta prop- sis functions are defined by a knot vector and the degree of the
erty), indicating that the boundary basis functions are continuously spline. A knot vector is a non-decreasing sequences of coordinates
extended into the interior. Other properties of Ti(n) include partition in the parametric space, written as U = {u1, u2, . . . , um+p+1}. The basis
of unity, positiveness, linear independence, and C1-continuity [6]. functions are defined recursively:
Next, assume that a piecewise smooth closed planar curve C is
u  ui uiþpþ1  u
parameterized, either exactly or approximately, in the Bezier form: Si;p ¼ Si;p1 þ Siþ1;p1 ; ð9Þ
uiþp  ui uiþpþ1  uiþ1
X
n
xðtÞ ¼ Ui ðtÞPi ; x 2 C; ð5Þ where for zero degree:
i¼1 
1 if ui 6 u < uiþ1 ;
where U(t) are the foregoing circular Bezier basis. Eq. (5) defines a Si;0 ¼ ð10Þ
0 otherwise:
mapping f: S ´ C, assumed to be invertible. Let P be the simply-con-
nected domain enclosed by C. Making use of the Harmonic coordi- B-splines and NURBS have been extensively covered in the literature,
nates, we can extend the mapping f into the domain, viz.: see e.g. [14]. Useful properties of B-spline basis are listed therein.
J. Lu, X. Zhou / Comput. Methods Appl. Mech. Engrg. 200 (2011) 233–241 235

Fig. 1. 3-Point, 4-point, 6-point, and 8-point circle and control polygons.

Fig. 2. A basis function in the 8-point disk. The dots on the circle are the physical
knots.

Fig. 4. An elliptical ring with control points.

to a general cylinder. The parametric domain is a straight cylinder


with unit circular cross section and a NURBS knot span
[u1, um+p+1]. If the Jacobian of (11) has the same sign throughout
the domain, the mapping is one-to-one.
Fig. 4 illustrates an elliptical ring represented as a tensor-prod-
uct rod. The cross section is an ellipse by four control points; the
centerline is a circle, defined by four 2nd-degree rational Bezier
arcs of equal length. There are a total of 32 control points. The cen-
terline is defined over the knot vector:
f0; 0; 0; 1; 1; 2; 2; 3; 3; 4; 4; 4g:
The surface knot lines (namely, iso-lines of u = ui on the lateral sur-
face) are also plotted.
Fig. 3. A flower-shape concave section generated by 16 control points.
2.3.1. Properties of the basis functions
The basis functions (12) have the following properties:
2.3. Tensor-product rod
P
1. Partition of unity: a N a ðn; uÞ ¼ 1 8ðn; uÞ 2 X.
Let us assume that the cross sections of a rod-like body are P
2. Linear independence: a ba N a ðn; uÞ ¼ 0 () ba ¼ 0 8ðn; uÞ 2 X.
parameterized by the harmonic mapping (6) with n control points, 1
3. Smoothness: Na is C in the subdomain D  (ui, ui+1), and is
and that the centerline is described as a NURBS curve with m con- (p-k)-times continuously differentiable with respect to u at a
trol points. To parameterize the rod, each centerline control point centerline knot, where k is the multiplicity of that knot.
is replaced by a cross section. Combining the cross sectional and 4. Positiveness: Na(n, u) P 0 " (n, u) 2 X.
axial representations gives the following tensor-product rod:
X
n X
m X
n X
m
  The first property ensures the linear consistency in an isopara-
x¼ T i ðn1 ; n2 ÞRj ðuÞPij :¼ Nij n1 ; n2 ; u Pij ; ð11Þ metric setting. As a consequence, the mapping (11) is affine invari-
i¼1 j¼1 i¼1 j¼1 ant in the sense that applying an affine transformation to the
where control points results in the same affine transformation to the en-
    tire volume.
Nij n1 ; n2 ; u ¼ T i n1 ; n2 Rj ðuÞ: ð12Þ
P Remark 2.3. On the lateral boundary the tensor product (11)
We can write Eq. (11) simply as x ¼ a N a Pa . This equation defines
essentially defines a swept surface for which there are well-
a mapping that mappings the cylindrical domain
  developed techniques for determining the parameters. In general,
X ¼ D  u1 ; umþpþ1 if the control points Pi of the generator and the NURBS parameters of
n o the centerline are known, the volume control points can be
¼ n1  n2  u : ðn1 Þ2 þ ðn2 Þ2 6 1; u1 6 u 6 umþpþ1
populated according to:
236 J. Lu, X. Zhou / Comput. Methods Appl. Mech. Engrg. 200 (2011) 233–241

Pij ¼ Cj þ Rj Sj Pi ; ð13Þ nates are employed. If one places r quadrature points in the radius
direction, s in the h direction, and t in the longitudinal direction,
where Cj are the control points of the centerline, Rj and Sj are rota-
the integration of a function F is computed according to:
tion and stretching tensors. In a forward design, the parameters Pi,
Cj, R, S are user-selected to produce the desired geometry. In an in- Z " #
X
n X
r X
s X
t
 
verse design, these parameters need to be determined by fitting. See Fdv  F nijk ; ul r ij wrj whk wzl ; ð14Þ
[14,24] for further discussions. X i¼1 j¼1 k¼1 l¼1

where (nijk, ul) are the position of a Gauss point, rij is the correspond-
Remark 2.4. In deformable modeling the representation (11) ing radius, wrj, whk, and wzl are the weights in r-, h-, and z-coordinate
admits a full three-dimensional deformation. In particular, warp- respectively. This quadrature rule is denoted as n  r  s  t, where
ing is allowed. the first number signifies the circumferential divisions.
We utilized the following configuration:
3. Cylindrical element r ¼ 2; s ¼ 1; t ¼ p þ 1:

The cylindrical mapping (11) furnishes a geometric basis for There are p + 1 layers of quadrature points placed longitudinally.
analysis in the isoparametric setting. In this approach, the refer- Each layer contains two rings of quadratures placed uniformly in
ence (undeformed) geometry and the current (deformed) geome- the circumferential direction, m point per ring. Numerical tests
try are described by the same basis functions; the latter is (not reported) indicated that this scheme appears to be suffi-
obtained by displacing the control points. The derivation of dis- ciently accurate and preserve the rank of the element stiffness
crete equations follows that of the standard isoparametric element, matrix.
and the details are omitted here.
NURBS geometry has an intrinsic mesh structure. A NURBS 4. Numerical Examples
curve consists of several rational Bezier segments that are formally
equivalent to elements. Each segment (i.e., element) corresponds 4.1. Patch test: uniaxial stretching of a cylindrical element
to a non-empty knot span [ui, ui+1]. The basis function is C1 within
a segment but suffers a loss of continuity at a knot point (e.g. the A finite strain patch test is conducted. The tested element is the
element boundaries). Exploiting this feature, it is natural to divide 12-node element (Fig. 5(a)) of unit radius and unit height. The
a tensor-product rod into several elements. Each element is a sub- material is nonlinear elastic with a neo-Hookean energy function:
cylinder corresponding to a non-empty knot span of the centerline.
l j
A typical element would have the parametric domain: WðCÞ ¼ ðI1  2 log J  3Þ þ ðlog JÞ2 ; ð15Þ
n o 2 2
Xi ¼ n1  n2  uÞ : ðn1 Þ2 þ ðn2 Þ2 6 1; ui 6 u 6 uiþ1 : where I1 = tr (C) is the first principal invariant of the Green-Lagrang-
ian deformation tensor C = FTF, (l, j) are the material constants, and
A physical element is the image of a parametric element, which is a J is the determinant of the deformation gradient F. In small strain
cylinder of radius 1 and length ui+1  ui. Depending on the cross sec- limit, l ¼ 2ð1þ
E Em
mÞ, j ¼ ð1þmÞð12mÞ, where E and m are Young’s modulus
tion and centerline parameters, an element can have variable num-
and Poisson’s ratio, respectively. In this example, the material
ber of control points. If one utilizes the n-point reduced Bezier form
parameters are taken to be E = 106 and m = 0.3.
for the cross section and a degree-p centerline, then the element
Lateral deformation tests were conducted previously [6] and
contains n  (p + 1) control points. Fig. 5 shows some examples
hence, axial loading is considered. The element is loaded axially
of parametric elements with the centerline domain [1, 1].
by displacement control points. The four control points at the low-
er end are fixed in the axial direction, and the control points at the
3.0.2. Quadrature rule
opposite end are displaced 0.5 unit axially. A minimal set of con-
straints just enough to eliminate rigid body motions is placed in
To compute the weak form one needs to integrate functions
the x- and y-directions. The axial displacements of the mid four
over an element. The computation boils down to an integration
control points are funded to be 0.25, and the axial strain
over the cylindrical parametric domain. We experimented some 1
ðC ZZ  1) has a uniform value 0.625 (Fig. 6). The x- and y-displace-
quadrature rules and the one which was found adequate is de- 2
ments of all control points agree with the analytical solution with-
scribed below.
in the machine precision.
The cylindrical domain is divided into n-angular slices of 2np
each, where n is the number or circumferential control points.
Within each slice, Gauss quadratures in cylindrical-polar coordi-

Fig. 5. Examples of parametric element: left: 2nd-degree 12-node element, right: Fig. 6. Uniaxial stretch of a cylindrical element: (a) deformed element; (b) strain
2nd-degree 24-node element. distribution.
J. Lu, X. Zhou / Comput. Methods Appl. Mech. Engrg. 200 (2011) 233–241 237

Fig. 7. Helical spring: (a) NURBS mesh with 4 elements per turn, (b) FEM model.

Table 1
Reaction force versus refinement (FEM result: 6.01). Row: number of elements per
turn in the longitudinal direction. Column: number of control points along the
circumference.

4 8 12 16 20 24 Fig. 9. Turbine blade profile.

4 8.23403 6.30259 6.15576 6.12637 6.11772 6.11448


8 8.21822 6.29145 6.14359 6.11389 6.10511 6.10181
16 8.21819 6.29142 6.14356 6.11385 6.10507 6.10177 we introduced a finite element model (Fig. 7(b)) that has in total
7081 nodes (6048 elements). The 8-node enhanced strained ele-
ment [25] is used. The finite element model gives a total reaction
force of 6.0100 in the z-direction. A convergence study is con-
4.2. Helical spring
ducted by increasing both the circumferential control points the
longitudinal knot divisions. The results for various mesh configu-
To quantitatively assess the cylindrical element, we simulated a
rations are presented in Table 1. Evidently, the number of cir-
helix spring under axial stretching and made a comparison with fi-
cumferential control points has little effect on the solution,
nite element analysis. The spring, as shown in Fig. 7, has three full
although small improvement can be observed. In contrast,
coils with a mean coil radius of 6 and a pitch of 2.60258; and the
increasing the longitudinal divisions leads to significant improve-
wire radius is 1. The centerline (helix) is a degree two NURBS
ments. For example, for the case with 4 circumferential control
curve, initially constructed from the knot vector:
points and 24 elements per turn, the result is 6.11448 (about
f0;0;0;1;1;2;2;3;3;4;4;5;5;6;6;7;7;8;8;9;9;10;10;11;11;12;12;12g; 1.6% difference to the FEM value). Fig. 8 shows the corresponding
von Mises stress distribution and a comparison with the finite
and nine control point per coil, with weights: element stress result.


1 1 1
1; pffiffiffi ; 1; pffiffiffi ; . . . ; 1; pffiffiffi ; 1 : 4.3. Turbine blade modal analysis
2 2 2
Each coil contains 4 cylindrical elements. Due to the repeated knot The blades of small wind turbines are often a solid body which,
values, the NURBS basis functions are C0 continuous across the ele- within the proposed paradigm, may be described as a rod. Fig. 9
ment boundaries. shows a cross section profile described by 16 control points. The
The same neo-Hookean model is used for the simulation. The blade in Fig. 10 is constructed by extruding a cross section along
lower end of the helix is fixed while the upper end is pulled up- the longitudinal axis while twisting the pitch angle and reducing
wards with a displacement of 2.0. The initial NURBS model con- the cross section dimensions at constant rates. The initial NURBS
tains four 12-node element per turn (Fig. 7(a)). For comparison, model contains a single element containing 16  3 = 48 control

50.0
50.0
45.0
45.0
40.0 40.0
35.0
35.0
30.0
30.0
25.0
25.0
20.0
20.0
15.0
15.0
10.0
10.0

5.0 5.0

Fig. 8. The von Mises stress: (a) cylindrical element, 24 elements per turn; (b) finite element. The degree-of-freedom numbers in these models are 1740 and 21243,
respectively.
238 J. Lu, X. Zhou / Comput. Methods Appl. Mech. Engrg. 200 (2011) 233–241

Fig. 10. Turbine blade, two views.

One is the line-of-action model which describes a muscle as a 3D


curve. This type of model is typically used for muscle force analysis
[26]. The other is the full scale 3D finite element model which in
general is used to study isolated muscles with emphasis on muscle
stress [27,28]. The challenges in modeling skeletal muscles lie in
capturing the complicated behavior and geometry. The line of ac-
tion model can only capture the gross behavior but incapable of
describing the geometry and muscle stress. Full scale finite ele-
ment simulations could be overly expensive especially when deal-
Fig. 11. The 10-element NURBS mesh of the blade. ing with muscle groups. When efficiency is of primary importance,
it is critical to balance physical realism and computational effi-
ciency. In [29,30], the present authors proposed to use NURBS-
points. The parameters are recorded in Appendix A. The model is based isogeometric approach for muscle analysis. This method
subsequently refined by knot insertion in the longitudinal direc- opens a pathway for a high-fidelity and high efficiency modeling
tion. A NURBS mesh containing 10 elements is shown in Fig. 11. of muscles.
The material is linear elastic, with E = 3  106, m = 0.3, and density Muscle tissue has a highly complex material behavior: it is ac-
q = 40. tive, nonlinear, incompressible, anisotropic, and hyperelastic
Table 2 presents the natural frequencies of the first ten modes, [31,27,28]. The microstructure of muscle is determined by the
computed from the rod model and a comparative finite element arrangement of the contractile elements (muscle fibers) and pas-
model. The finite element mesh contains 1035 nodes (686 8-node sive background tissue within a muscle. At the continuum scale,
elements). In comparison, the finest NURBS mesh (30 elements) muscles are modeled as hyperelastic, anisotropic material embed-
contains 32  16 = 512 control points. Clearly this level of mesh ded with active contractile fibers. In this study, the muscle is as-
density is not necessary as converged results are obtained with sumed to be hyperelastic solid with the following strain energy
20 elements (352 control points). The first five frequencies are in function:
good agreement with the finite element results. The finite element
analysis utilized the enhanced strain element; the pure displace- WðCÞ ¼ W matrix ðCÞ þ W fiber ðCÞ: ð16Þ
ment 8-node element is overly stiff for this problem due to shear
locking. The modal shapes for the first ten modes are depicted in The first term is the strain energy associated with the passive
Fig. 12. ground substance and the second term represents the active and
Fig. 13 shows the convergence behavior for the first three passive muscle fiber strain energy. Following [28], the muscle fiber
modes, taking the 30-element model as the reference. The strain energy assumes the form:
  vertical
axis corresponds to the logarithmic errors log xx  1 . A cubic
ref W fiber ðCÞ ¼ W act ðk; aÞ þ W pass ðkÞ; ð17Þ
rate of convergence is observed.
where Wact(k, a) and Wpass(k) are, respectively, the active and pas-
5. Skeletal muscle modeling sive strain energy of the muscle fiber. In addition to a dependence
on the fiber stretch k, the active response is also regulated by a sca-
Slender muscles present a good example of soft biological or- lar parameter a representing the level of activation (which is con-
gans that may be effectively described by cylindrical elements. trolled by neural excitation). The derivative of Wfiber(k, a) with
Existing skeletal muscle models typically fall into two categories. respect to the stretch k is defined as

Table 2
The first ten frequencies. N: number of NURBS elements.

Modes Frequencies
N=5 N = 10 N = 15 N = 20 N = 25 N = 30 FEM
1 0.19136 0.18857 0.18812 0.18796 0.187874 0.187824 0.1909
2 0.95579 0.79195 0.77924 0.77668 0.775798 0.775383 0.8018
3 1.44902 1.43379 1.43130 1.43032 1.42979 1.42945 1.3674
4 2.49828 2.06955 1.96951 1.95178 1.94657 1.94448 2.0630
5 3.46974 2.46023 2.45053 2.44705 2.44552 2.44473 2.2571
6 5.30767 4.20237 3.76425 3.69110 3.67126 3.66391 4.0258
7 5.63743 5.09790 5.06690 5.05624 5.05173 5.04947 4.4984
8 9.64319 5.53363 5.51095 5.50019 5.49542 5.49304 5.3152
9 10.4882 7.63507 6.24509 6.02659 5.97116 5.9518 6.8215
10 10.7239 8.16530 8.06399 8.03176 8.01842 8.01226 7.2485
J. Lu, X. Zhou / Comput. Methods Appl. Mech. Engrg. 200 (2011) 233–241 239

−1
ω1
−2 ω2
ω3

−3

1
−4
3

Log(error)
−5

−6

−7

−8

−9
1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2
Log(N)

Fig. 13. Rate of convergence for the first three frequencies. N: number of NURBS
elements.

Fig. 14. Biceps initial configuration.

where rmax is the maximum active stress that can be generated by


the muscle fiber, m1 and m2 are the coefficients for the passive
property of muscle fiber. With these terms, the well-known nonlin-
ear stress–strain relationship of muscle fibers [32,33] can be mod-
Fig. 12. Modal shapes for the first ten modes.
eled at the continuum level. Further derivations of the stress
tensors are ignored here. More details can be found in [30].
For this example, the stain energy function of the ground (ma-
oW act ðk; aÞ h i trix) material, Wmatrix, takes the same form as neo-Hookean mate-
¼ armax 1  4ð1  kÞ2 ; ð18Þ rial with E = 30 N/cm2, m = 0.49. Other muscle material parameters
ok
are: m1 = 35, m2 = 0.5, and rmax = 30 N/cm2. To emulate the mus-
and
( cle–tendon complex, the first and the last element are modeled
oW pass ðkÞ m1 kðem2 ðk
2
1Þ
 1Þ; k > 1; as tendon and the mid three elements are modeled with active
¼ ð19Þ muscle material. The tendon is modeled by the ground material
ok 0; k 6 1;
alone but ten times stiffer.
240 J. Lu, X. Zhou / Comput. Methods Appl. Mech. Engrg. 200 (2011) 233–241

ber direction is (0, 1, 0), which is approximately aligned with the


muscle action line (which links the two end surfaces). The two
end surfaces of the muscle solid are fixed to keep the muscle length
unchanged. The contact condition between the two branches is de-
4 30 scribed using the isogeometric contact formulation [34]. Note here
25
we only handle the contact between the muscle elements because
3
initially the two lower tendons largely overlap each other (we in-
20
tend to keep them so) and the two upper tendons are far apart.
2 15
We employ explicit dynamics method for time integration and
10 the time step is 2.5  105 second. In the simulation, we gradually
1
5
(linearly) increase the muscle activation up to the maximum level
(i.e. a = 1) in the first 100 steps (<0.25 ms) and then keep it con-
0 0
stant at the maximum level. Fig. 15(a)–(c) present the deformed
muscle shapes and von Mises stresses at 5, 10, and 15 ms, respec-
tively. The muscle belly bulges up due to fiber contraction and con-
sequently the stress increases initially. The maximum stress
happens at the two tendon sections due to their small cross section
area. We utilize significant numerical damping so that muscle
reaches equilibrium (almost no further deformation) at 250 ms.
The final state is shown in Fig. 16(a). As can be observed, the stress
level in overall is slightly lower than that at 0.25 ms. This is be-
cause the muscle fiber has a higher contraction strength at low
30 contraction strains. As the muscle fiber contracts shortens, its
capacity to generating force decreases although the activation is
25
kept at the same level [30].
20
For comparison, we also simulated the same with FEM, for
15 which a tetrahedron mesh of the muscle with 7413 nodes
10
(25315 elements) is used. As in the cylindrical element approach,
contact is defined between the two muscle bellies. The equilibrium
5
state (again at 250 ms) from the FEM simulation is shown in
0
Fig. 16(b). By comparison, it can be seen that stress distributions
are in good qualitative agreement with each other.

6. Concluding remarks

We introduced a new approach for analyzing rod-like struc-


Fig. 15. The von Mises stress (plotted on the deformed configuration) at (a) 5 ms;
tures. The method is marked by the utility of CAD and CG geome-
(b) 10 ms; and (c) 15 ms.
tries; a rod is represented as a tensor-product volume using
harmonic coordinates and NURBS. Reduction of dimension is
achieved at the geometry level, by using only surface control
points. This representation preserves the geometry of the rod and
smoothness. The model eliminates the conventional assumption
of rigid cross section and hence, allows for the consideration of lat-
eral deformations and lateral contact. This feature is significant for
30
analyzing slender soft structures which normally undergo large
25 lateral deformations. In the traditional paradigm, structures as
20 such are treated as a three-dimensional continuum.
15
Appendix A. Parameters of the turbine blade
10

5
The generator (the root cross section) are parameterized by the
0 following control points:
2 3
ð1:; 0:0708625; 0:Þ; ð0:847759; 0:177017; 0:Þ;
6 ð0:566454; 0:227835; 0:Þ; ð0:198912; 0:225; 0:Þ; 7
6 7
6 7
6 ð0:198912; 0:18125; 0:Þ; ð0:566454; 0:116567; 0:Þ; 7
6 7
6 ð0:847759; 0:0531051; 0:Þ; ð1:; 0:00994562; 0:Þ; 7
6 7
6 7:
6 ð1:; 0:00397825; 0:Þ; ð0:847759; 0:00566454; 0:Þ; 7
Fig. 16. The deformed configuration and von Mises stress at equilibrium (250 ms): 6 7
(a) cylindrical element, (b) FEM. 6 ð0:566454; 0:; 0:Þ; ð0:198912; 0:01; 0:Þ; 7
6 7
6 7
4 ð0:198912; 0:020:Þ; ð0:566454; 0:0254328; 0:Þ; 5
ð0:847759; 0:0226582; 0:Þ; ð1:  0:00994562; 0:Þ
The biceps in Fig. 14 are described by cylindrical elements. Each
branch contains 5 elements, with 8 control points for the cross sec- For the centerline, the initial NURBS parameters are
tion and a quadratic NURBS centerline. There are a total of
u ¼ f0; 0; 0; 1; 1; 1g; p ¼ 2; w ¼ f1; 1; 1g;
7  8 = 56 control points in each branch. We assume the muscle fi-
J. Lu, X. Zhou / Comput. Methods Appl. Mech. Engrg. 200 (2011) 233–241 241

and control points: [13] B. Nadler, M.B. Rubin, A new 3-D finite element for nonlinear elasticity using
2 3 the theory of a Cosserat point, Int. J. Solids Struct. 40 (2003) 4585–4614.
1 0 0 [14] L.A. Piegl, W. Tiller, The NURBS Book, Springer, Berlin, 1995.
  6 7 [15] T.J.R. Hughes, J.A. Cottrell, Y. Bazilevs, Isogeometric analysis: CAD, finite
Cj ¼ 4 1 0 4:15 5: elements, NURBS, exact geometry and mesh refinement, Comput. Meth. Appl.
1 0 8:30 Mech. Engrg. 194 (2005) 4135–4195.
[16] Y. Bazilevs, L. Beirao de Veiga, J.A. Cottrell, T.J.R. Hughes, G. Sangalli,
The control points of the blade are computed according to Eq. Isogeometric analysis: approximation, stability and error estimates for h-
refined meshes, Math. Model Meth. Appl. Sci. 16 (2006) 1031–1090.
(13). For the root layer: [17] J.A. Cottrell, A. Reali, Y. Bazilevs, T.J.R. Hughes, Isogeometric analysis of
structural vibrations, Comput. Meth. Appl. Mech. Engrg. 195 (2006) 5257–
Pi1 ¼ Pi ; i ¼ 1; . . . ; 16: 5296.
[18] J.A. Cottrell, A. Reali, T.J.R. Hughes, Studies of refinement and continuity in
For the middle layer:
isogeometric structural analysis, Comput. Meth. Appl. Mech. Engrg. 196 (2007)
4160–4183.
Pi2 ¼ C2 þ R2 S2 ðPi  C2 Þ; i ¼ 1; . . . ; 16;
2   3 23 3 [19] T. Elguedj, Y. Bazilevs, V.M. Calo, T.J.R. Hughes, B and F projection methods for
cos 2b  sin 2b 0 4
0 0 nearly incompressible linear and non-linear elasticity and plasticity using
6   7 6 7 higher-order NURBS elements, Comput. Meth. Appl. Mech. Engrg. 197 (2008)
R2 ¼ 4 sin 2b cos 2b 0 5; S2 ¼ 4 0 34 0 5: 2732–2762.
[20] Y. Bazilevs, V.M. Calo, J.A. Cottrell, J.A. Evans, T.J.R. Hughes, S. Lipton, M.A.
0 0 1 0 0 1
Scott, T.W. Sederberg, Isogeometric analysis using T-splines, Comput. Meth.
Appl. Mech. Engrg. 199 (2010) 229–263.
For the tip layer: [21] T.J.R. Hughes, A. Reali, G. Sangalli, Efficient quadrature for NURBS-based
isogeometric analysis, Comput. Meth. Appl. Mech. Engrg. 199 (2010) 301–313.
Pi3 ¼ C3 þ R3 S3 ðPi  C3 Þ; i ¼ 1; . . . ; 16; [22] C. Blanc, C. Schlick, Accurate parameterization of conics by NURBS, IEEE
2 3 21 3
cos b  sin b 0 2
0 0 Comput. Graph. Appl. 16 (1996) 64–71.
6 7 6 7 [23] F. John, Partial Differential Equations, third ed., Springer, New York, 1978.
R3 ¼ 4 sin b cos b 0 5; S3 ¼ 4 0 12 0 5; [24] M. Aigner, C. Heinrich, B. Juttler, E. Pilgerstorfer, B. Simeon, A.V. Vuong, Swept
0 0 1 0 0 1 volume parameterization for isogeometric analysis, in: Mathematics of
Surfaces XIII, Lecture Notes in Computer Science, Springer, Berlin,
where b = 0.268079 in radian. Heidelberg, 2009, pp. 19–44.
[25] J.C. Simo, F. Armero, R.L. Taylor, Improved versions of assumed enhanced strain
tri-linear elements for 3D finite deformations, Comput. Meth. Appl. Mech.
References Engrg. 110 (1993) 359–386.
[26] S.L. Delp, J.P. Loan, A computational framework for simulation and analysis of
[1] A.E.H. Love, A Treatise on the Mathematical Theory of Elasticity, fourth ed., human and animal movement, IEEE Comput. Sci. Engrg. 2 (2000)
Dover, New York, 1944. 46–55.
[2] S.S. Antman, The theory of rods, Handbuch de Physik, vol. Via/2, Springer, [27] R. Lemors, M. Epstein, W. Herzog, B. Wyvill, Realistic skeletal muscle
Berlin, 1972, pp. 641–703. deformation using finite element analysis, in: Proceedings of the 14th
[3] J.C. Simo, A finite strain beam formulation. I. The three-dimensional dynamical Brazilian Symposium on Computer Graphics and Image Processing, IEEE
problem, Comput. Meth. Appl. Mech. Engrg. 49 (1985) 55–77. Computer Society Press, 2001, pp. 192–199.
[4] J.C. Simo, L. Vu-Quoc, A three-dimensional finite-strain rod model. II. [28] C.W.J. Oomens, M. Maenhout, C.H.G.A. van Oijen, M.R. Drost, F.P.T. Baaijens,
Computational aspects, Comput. Meth. Appl. Mech. Engrg. 58 (1986) 79–116. Finite element modelling of contracting skeletal muscle, Philos. Trans. R. Soc.
[5] M.A. Crisfield, Nonlienar Finite Element Analysis of Solid and Structures, vol. 2, Lond. B 358 (2003) 1453–1460.
John Wiley and Sons, New York, 2003. [29] X. Zhou, J. Lu, NURBS-based Galerkin method and application to skeletal
[6] J. Lu, Circular element: isogeometric elements of smooth boundary, Comput. muscle modeling, in: Proceedings of the 2005 ACM Symposium on Solid and
Meth. Appl. Mech. Engrg. 198 (2009) 2391–2402. Physical Modeling, SPM’05, ACM Press, New York, 2005, pp. 71–78.
[7] A.E. Green, P.M. Naghdi, M.L. Wenner, On the theory of rods I: derivation from [30] X. Zhou, J. Lu, Biomechanical analysis of skeletal muscle in an interactive
the three dimensional equations, Proc. R. Soc. Lond. A 337 (1974) 451–483. digital human system, SAE Trans. J. Passenger Cars Mech. Syst. 114 (2005)
[8] A.E. Green, P.M. Naghdi, M.L. Wenner, On the theory of rods II: development by 2921–2929.
direct approach, Proc. R. Soc. Lond. A 337 (1974) 485–504. [31] W. Herzog (Ed.), Skeletal Muscle Mechanics: From Mechanisms to Function,
[9] P.M. Naghdi, M.B. Rubin, Constrained theories of rods, J. Elasticity 14 (1984) John Wiley and Sons, Chichester, 2000.
343–361. [32] F.E. Zajac, E.L. Topp, P.J. Stevenson, A dimensionless musculotendon model, in:
[10] P.M. Naghdi, M.B. Rubin, On the significance of normal cross sectional Proceedings IEEE Engineering in Medicine and Biology, 1986.
extension in beam theory with application to contact problem, Int. J. Solids [33] F.E. Zajac, Muscle and tendon: properties, models, scaling, and application to
Struct. 25 (1989) 249–265. biomechanics and motor control, Crit. Rev. Biomed. Engrg. 17 (1989)
[11] O.M. O’Reilly, On the constitutive relations for elastic rods, Int. J. Solids Struct. 359–411.
35 (1998) 1009–1024. [34] J. Lu, Isogeometric contact analysis: geometric basis and formulation for
[12] M.B. Rubin, Numerical solution procedures for nonlinear elastic rods using the frictionless contact. Comput. Meth. Appl. Mech. Engrg., submitted for
theory of a Cosserat point, Int. J. Solids Struct. 38 (2001) 4395–4437. publication, <http://css.engineering.uiowa.edu/jialu/html/text-contact.pdf>.

You might also like