Comparison between Bézier Extraction and Associated Bézier Elements
in Eigenvalue Problems
CHRISTOPHER G. PROVATIDIS
School of Mechanical Engineering,
National Technical University of Athens,
9, Iroon Polytechniou str, 15780 Zografou,
GREECE
Abstract: - This paper shows that the control points which are implicitly encountered in the Bézier extraction
during isogeometric analysis can be explicitly used to form zier elements of 𝐶-continuity in several ways,
thus eventually leading to a superior accuracy and performance than the 𝐶-continuity. The study is reduced to
the eigenvalue extraction in problems governed by the Helmholtz equation. Analysis is performed in
conjunction with piecewise cubic interpolation for three benchmark tests in one and two dimensions. In the
latter case a rectangular and a circular acoustical cavity under Neumann boundary conditions are analyzed.
Several computational details are also discussed.
Key-Words: - Bézier extraction; isogeometric analysis; NURBS; finite elements; Helmholtz equation
Received: March 28, 2022. Revised: November 24, 2022. Accepted: December 21, 2022. Published: December 31, 2022.
1 Introduction
Within the context of the finite element method
(FEM), the idea of using the same set of basis
functions for the geometry and the analysis is very
old. The practical need of this concept is the
treatment of curvilinear domains where the Jacobian
matrix has to be found at the integration points
aiming at estimating the stiffness (static problems)
as well as the mass matrix (time-dependent
problems). The first report referring to the so-called
isoparametric element is due to Taig, [1]. A later
public paper is due to Irons, [2], while a detailed
description may be found in classical FEM
textbooks (see [3], [4]).
The above idea, i.e., the use of the same
functional set for both the geometric model and the
analysis, was later extended using the Coons
interpolation method (see [5], [6]). The difference of
the latter with the previous standard FEM methods
is that, instead of using a known functional set such
as a power series expansion or a sinusoidal series,
global shape functions are extracted from a
computer-aided geometric design (CAGD)
interpolation formula. These shape functions span
an entire patch of the domain which, in general, may
be also decomposed into a number of smaller
curvilinear patches.
As has been documented in classical CAGD
textbooks such as, [7], the abovementioned Coons
interpolation formula, due to Steven Coons (1964-
1967), is almost the first one which was ever
developed in the theory of CAGD. Moreover, in
chronological order, the second important formula
was due to William Gordon (1969), a third
interpolation was due to Pierre Bézier (1970), a
fourth interpolation was the ‘B-spline’ due to Curry-
Schoenberg formulation (1966) that was later
treated more efficiently thanks to an efficient
recursive scheme by Cox-deBoor-Mansfield (1972),
the fifth was NURBS due to Versprille (1975) but
was popularized twenty years later mainly by the
monograph of Piegl and Tiller containing pseudo-
codes, [8]. The sixth cornerstone CAGD-
interpolation is T-splines, [9], while it is anticipated
that novel CAGD-based interpolations will follow to
keep busy the next generations of FEM-researchers.
The author has not included a lot of other significant
interpolations (e.g. alphabetically: Bajaj-, Barnhill-,
DeRose-, Gregory-patches, etc.) because he had to
restrict himself to those advertised in the most
known commercial CAD/CAM/CAE software
packages. The research on splines, as a means to
solve partial differential equations (PDEs) with
given boundary conditions, still continues to this
date ( [10], [11], [13]).
Based on transfinite Coons-Gordon interpolation
formulas a number of papers have appeared for
either static or dynamic analysis ([14] and [15]),
while [16] is a monograph including the summary of
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
605
Volume 17, 2022
almost sixty papers (for the period 1989 to 2017)
which was published much later in 2019.
Based on B-spline interpolation a number of
papers on the numerical solution of PDEs have
appeared near 2000 (see [17], [18], [19]), while a
relevant monograph is [20].
Regarding the application of non-uniform
rational B-splines (NURBS) in engineering analysis,
the first relevant studies were reported in 1993 and
1994 (see [21] and [22]). NURBS were revived in
2005 under the title ‘NURBS-based isogeometric
analysis’ (IGA). This was done mainly for product
shape design, [23]. In the same year (2005), [24], an
extensive paper dealt with a broader spectrum of
engineering applications for which all the well-
known flexibilities and advantages of previously
developed CAGD algorithms such as knot insertion,
degree elevation, etc were utilized. A relevant
textbook, published in 2009, is [25].
In its original implementation, IGA requires
substantial computer time in order to calculate the
basis functions at the integration points thus to
estimate the stiffness matrix, despite the local
support of the basis and the fact that the efficient de
Boor-Cox recursive algorithm was used (see [26],
[27]). To reduce the computer effort, the Bézier
extraction technique has been proposed, [28]. In
brief, the main advantage is that Bézier extraction,
compared to the original implementation of IGA,
[25], is such that apart from the coefficients of the
extraction operator (matrix Ce) of the e-th’
NURBS-element within a patch, the basis functions
are identical for all elements in the mesh as it is the
case for classical finite elements. Therefore, there is
no need to implement B-spline basis function
evaluation routines which (as previously said) are
costly from a numerical point of view.
Within the context of (tensor-product) NURBS
approximation of degree p (in conjunction with
control points P), after the computation of the
extraction operator Ce (for definitions, see [28]) the
updated set of control points Q (associated
Bernstein-Bézier polynomials) can be easily
calculated by the linear formula

,
e
QCP
where
the superscript T stands for the transpose of the
matrix Ce. As a result, if the aforementioned control
points Q are properly grouped, they can build a set
of Bézier elements of degree p with C0 inter-
element-continuity.
In this paper it will be shown that the above-
mentioned control points Q, which again are
implicitly encountered in the abovementioned
Bézier extraction (MODEL-1), can be explicitly
used to form Bézier elements of 𝐶-continuity
(henceforth MODEL-2) with superior accuracy.
2 The Main Idea
The numerical method of ‘Bézier extraction’ is
based on the insertion of additional knots at all the
inner knots of the patch until their multiplicity
becomes equal to the polynomial degree 𝑝. Then a
linear relationship is derived between the NURBS
basis functions N and the Bernstein polynomials B
according to the expression:
e
NCB
(1)
where e
C is the well known Bézier operator, [28].
Since after knot insertion the shape of the patch
remains unaltered, in form and parametrically, the
patch (,)

x is described by the condition:
(,)


xNPBQ
, (2)
where Q denotes the control points after the
abovementioned multiple knot insertion (those
involved in Bézier extraction). Substitution of Eq.
(1) into Eq. (2) and further drop out of the common
factor BT on the left of the second equality, leads to

e
CPQ
, (3)
or equivalently:
1
e
PC Q
(4)
In one-dimensional (1D)-problems, and for every
NURBS-element e’, P is of a column-vector of size
󰇛𝑝 1󰇜, Q is of a column-vector of size 󰇛𝑝 1󰇜,
whereas the matrix Ce is a matrix of size 󰇛𝑝 1󰇜
󰇛𝑝 1󰇜.
Implementing the concept of isogeometric
analysis (i.e., same basis functions for the analysis
as well), a similar relationship will hold in both
systems for the degrees of freedom (DOFs), a, as
well, thus the analogue of Eq. (4)will be:
1
PeQ
aCa
(5)
Regarding MODEL-1 (original IGA), for any
NURBS elemente let the corresponding stiffness
and mass matrices be e
K and e
M, respectively.
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
606
Volume 17, 2022
Below we shall derive the relevant matrices for
MODEL-2.
Actually, Eq. (5)depicts a linear relationship
between the ‘local’ degrees of freedom (DOFs), aP,
of MODEL-1 (B-spline or NURBS) and the ‘global’
DOFs, aQ, of MODEL-2, through the transformation
matrix
1
e
TC . This change of basis leads to the
well-known quadratic form (Kglobal = TT Klocal T)
thus the matrices of MODEL-2 can be written in
terms of the local matrices as follows:


11
eeee
Q

KCKC
(6)
and


11
eeee
Q

MCMC
. (7)
In addition to the general quadratic form of Eqs. (6)
and (7) which requires only matrix operations, three
more equivalent alternatives for MODEL-2 will be
discussed below.
To make the procedure as clear as possible, we
shall start with a typical 1D-problem and then
continue with a couple of 2D-problems.
2.1 One-dimensional problem
Problem: Consider the Helmholtz equation
222
()0ux ku
with wave-number kc
, in
the interval 󰇟0, 𝐿󰇠 with 𝐿3. The boundary
condition at the left end is of Dirichlet type (𝑢0
at 𝑥0) while at the right end is of Neumann type
(𝜕𝑢 𝜕𝑥
0 at 𝑥𝐿). Find the eigenvalues
implementing IGA for the polynomial degree 𝑝3
and three uniform B-spline elements.
Solution: The exact eigenvalues 𝜆𝜔
are
given by:
22 2
(2 1) (4 ), 1,2,
iicLi

 . (8)
2.1.1 Original IGA (MODEL-1)
First, B-splines analysis (MODEL-1) is performed
for the knot vector 𝛯 󰇝0,0,0,0,1,2,3,3,3,3󰇞 (i.e, for
𝑛3 uniform breakpoint spans), for polynomial
degree 𝑝3, which involves three B-spline
elements and six control points as shown in Fig.
1(a). In general, the position of the control points
may be arbitrary. However, if we allow a linear
isoparametric mapping, i.e. that 𝑥󰇛𝜉󰇜𝜉 with 󰇛0
𝜉3󰇜, then the aforementioned six involved
control points are found to be located at the
particular positions 𝑥 󰇝0,
,1,2,
,3󰇞 as shown
in Fig. 1(a). The aforementioned linear mapping is
not obligatory but permits the accurate
representation of ideally linear functions. Based on
these six control points, for each element ′𝑒󰆒 󰇛𝑒
1,2,3󰇜 the Jacobian will be a constant, because the
domain is a straight line. Wherever the control
points are, the six basis functions are those
illustrated in Fig. 2(a). Two of them span one-third,
also two of them span two-thirds and the final two
basis functions span the entire domain.
Elementary B-spline theory suggests that the
connectivity vector which allocates the global DOFs
to each of the abovementioned three B-spline
elements, will be IEN-1 = [(1,2,3,4), (2,3,4,5),
(3,4,5,6)], so as only 𝑝1 basis functions affect
each element. For each of these three elements the
local matrices are given by:
0
[] , 1,2,3,4
e
lj
ei
ij
N
Ndx i
xx


K (9)
and
2
0
1
[] , 1,2,3,4
e
l
e
ij i j
NNdx i
c

M. (10)
After matrix assembly, the boundary conditions
are imposed (deletion of the first raw and column in
both the mass and stiffness matrices due to the
Dirichlet boundary condition), and the eigenvalues
are calculated in terms of the system matrices, K
and M (after deletion) requiring that the determinant
vanishes:
det 0
i
KM , (11)
with 𝜆𝜔
denoting the 𝑖-th eigenvalue of the
problem.
Either the basis functions 𝑁 involved in Eqs. (9)
and (10) are computed in terms of usual B-splines
(e.g., using a standard function such as the spcol
in MATLAB) or in terms of Bernstein polynomials
using the Bézier operator e
C [see, Eq. (1)], the
obtained matrices are the same thus leading to the
numerical results shown in Fig. 3 (blue line). One
may observe that the computational error increases
with the serial number of the eigenmode.
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
607
Volume 17, 2022
Fig. 1: Arrangement of control points in (a) initial
B-spline and (b) after the multiple knot insertion.
Fig. 2: Basis functions for (a) initial B-spline and (b)
after the multiple knot insertion.
Fig. 3: Errors (in percent) of the calculated
eigenvalues (Model-1: 6 and Model-2: 10 control
points).
2.1.2 Bézier elements and similar (MODEL-2)
Increasing the multiplicity of the two inner knots (at
𝜉1 and 𝜉2) from ‘1’ to ‘3’ (note that 𝑝3),
the new control points increase from 6 (P) to 10 (Q)
and are uniformly arranged within the interval [0,3]
as shown in Fig. 1(b) while the corresponding basis
functions are shown in Fig. 2(b). In the latter figure,
one may observe that the new basis functions are
interpolatory at the ends of each element (which are
also ends of repeated knots) and span two elements
(one on the left and the other on the right of each
multiple inner knot). The new connectivity vector
will be: IEN-2 = [(1,2,3,4), (4,5,6,7), (7,8,9,10)].
Again, only 𝑝1 basis functions affect each
element.
At this point, four alternative procedures (all of
them labeled asMODEL-2’) are tested as follows:
2.1.2.1 Procedure (2a): de Boor B-splines
The first procedure for Model-2 consists of
practically using the same code as that of MODEL-1
but now setting the multiplicity of the (two) inner
knots equal to 𝑝3. For the produced updated
basis functions 𝑁,𝑖 1,…,10, we can use the
standard recursive (de Boor) formula, such as
spcol existing in MATLAB®. For each of the
three B-spline elements we use four (i.e., 𝑝1)
Gauss points and significant computer savings are
achieved when the local support is encountered
through the above-mentioned connectivity vector
IEN-2.
2.1.2.2 Procedure (2b): Bézier elements
The second procedure for Model-2 consists of using
the three cubic (4-node) Bézier elements of equal
size shown in Fig. 1(b) and Fig. 2(b). Each Bézier
element is made of four DOFs associated to four
Bernstein polynomials as basis functions. If one
does not wish to perform numerical Gauss
integration in conjunction with the aforementioned
Bernstein polynomials, the analytical formulas of
the matrices in each element are given in Appendix
A. The connectivity vector will be again the
aforementioned IEN-2.
2.1.2.3 Procedure (2c): Lagrange elements
The third procedure for Model-2 consists of using
the three classical cubic Lagrange elements of equal
size which are shown in Fig. 1(b). The analytical
formulas of the matrices in each element are given
in Appendix B. The same connectivity vector, i.e.,
IEN-2 will be applied.
2.1.2.4 Procedure (2d): Matrix transformation
The fourth and last alternative procedure for Model-
2 consists of blindly implementing Eqs. (6) and (7)
at each of the three B-spline elements shown in Fig.
1(a) that have been found in Model-1, and then
rearranging the produced quadratic form in the form
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
608
Volume 17, 2022
of enlarged matrices according to the connectivity
vector IEN-2. From the computational point of
view, we can use two loops as follows. The outer
loop refers to the three elements while the inner
loop refers to the (𝑝1) Gauss points in each
element. Between these two loops, the local
matrices (Ke, Me) are calculated and then
eQ
K as
well as
eQ
M are found according to Eqs. (6) and
(7).
It was found that three out of the four
abovementioned procedures (i.e., the Procedures 2a,
2b and 2d) lead to identical mass and stiffness
matrices. Not only that, but although the Procedure
2c gives different matrices than the other three
procedures, it eventually leads to the same
numerical results (calculated eigenvalues) with the
Procedures (2a, 2b and 2d). Therefore, the error of
the calculated eigenvalues for all these four
procedures, represented under the umbrella of
MODEL-2, is shown in Fig. 3 (red line).
2.1.3 Comparison of MODEL-1 with MODEL-2
One may observe that MODEL-1 [i.e., three B-
spline elements (6DOFs, 𝐶-continuity)] is less
accurate than MODEL-2 [i.e., a totality of three
Bézier elements (10 DOFs, 𝐶-continuity)]. One
reason for the superiority of the 𝐶-continuity
(MODEL-2) is probably the larger number of the
participating DOFs (10 versus 6 before the
imposition of the BCs, and 9 versus 5 after the
deletion of the restrained DOFs).
It is noted that in the above one-dimensional test
case it is very difficult to compare the CPU-times
because all of them are very small and the whole
comparison is very sensitive.
2.2 Two-dimensional problems
2.2.1 Rectangular cavity
Consider an acoustic cavity of size 𝑎𝑏2.5𝑚
1.1𝑚, with normalized wave speed 𝑐1𝑚 𝑠
, and
rigid walls (𝜕𝑢 𝜕𝑛
0). For the governing
Helmholtz equation 𝑢𝑘
𝑢0 (with 𝑘
𝜔𝑐
), calculate the lowest 15 eigenvalues 𝜆
𝜔
,𝑖 1,2,⋯,15. We recall that the exact
eigenvalues are given by the formula:
22
222 ,, 0,1,,
mn mn mn
cmn
ab









#(12)
Solution: The initial isogeometric model
(MODEL-1 of 𝐶-continuity) comprises 10 cubic B-
spline elements in a 52 setup, which includes 8
540 control points (see, Fig. 4(a)). Therefore,
each matrix of (Κ, Μ) comprises 40 rows and 40
columns, of which none is deleted when imposing
the Neumann-type BCs. The results of this model
are shown in Fig. 5: (blue line).
Fig. 4: Control points for (a) 𝐶- and (b) 𝐶-continuity, (c) detail of Bézier element No. 8.
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
609
Volume 17, 2022
After the zier decomposition, the totality of
the new control points are rearranged according to
Fig. 4(b), in which their number has increased (from
40) to 16 7 112. Now, each of the ten cubic
Bézier elements consists of 16 control points
arranged in a 44 setup. Here it happens that all
the extreme control points belong to the true
boundary of each element and they split each edge
in three equal pieces, as if we had to deal with
conventional rectangular finite elements. As
previously happened with the 1D-problem, four
equivalent procedures have been developed; three of
them are coincident (i.e., the procedures 2a, 2b, and
2d), one is equivalent (procedure 2c), and all of
them constitute MODEL-2’. The IEN-2 matrix can
be automatically computed in advance, based on the
formulas which are illustrated in Fig. 4(c), where the
indices (𝑖,𝑖
,𝑖
and 𝑖) denote the serial numbers of
the control points at the corners of each NURBS
element. In this convention, the Bézier elements are
numbered sequentially starting from the lower left
and increasing until the bottom edge is completed;
then we continue with the above-the-bottom layer
from left to right, and so on (see, the numbers inside
the small red circles of Fig. 3b). Again, Procedure-
2a involves the MATLAB’s function spcol in
conjunction with multiplicity 3 (or its equivalent
Bézier extraction), Procedure-2b includes the ten
tensor product cubic Bézier elements, Procedure-2c
includes ten conventional tensor product cubic
Lagrange elements, whereas the algebraic
Procedure-2d includes change of basis using Eqs.
(6) and (7). In all these four alternative procedures
(2a2d) for Model-2 the results were found to be
identical (see Fig. 5:, red line) and one may observe
that they outperform the Model-1.
Although MODEL-2 (based on 𝐶-continuity) is
superior to MODEL-1 (𝐶-continuity), we have to
admit that the former deals with 112 DOFs
compared to 40 DOFs existing in the later model.
Fig. 5: Errors of the calculated eigenvalues for the
rectangular cavity.
2.2.2 Circular cavity
A circular acoustic cavity of unit radius (
1a
)
under Neumann boundary conditions is studied. The
analytical solution is given by
( ) 0, 0,1, 2,
m
Jka m

#(13)
where
()
m
Jka
is the first derivative of the Bessel
function
()
m
Jka
of the first kind and order
m
and
kc
is the wavenumber. We wish to find the
lowest seventeen eigenvalues applying IGA.
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
610
Volume 17, 2022
(a) (b)
(c) (d)
Fig. 6: Progressive knot insertion in a circular cavity (a) quadratic Bézier with isolines, (b)
cubic Bézier, (c) cubic NURBS with 𝐶-continuity, (d) cubic NURBS with 𝐶-continuity.
Solution: The starting point is the well-known 9-
point rational Bézier tensor product (a 33 setup,
see Fig. 6a) for degree 𝑝2, knot vector 𝛯
󰇝0,0,0,1,1,1󰇞 󰇝0,0,0,1,1,1󰇞 and weights 𝒘
󰇝1,
√
,1󰇞󰇝1,
√
,1󰇞.
Then the degree is elevated to 𝑝3 and, as a
result, the new set of the sixteen control points are
arranged in a 44 setup, as shown in Fig. 6(b),
where the new knot vector per direction becomes
𝛯′ 󰇝0,0,0,0, 1,1,1,1󰇞 󰇝0,0,0,0, 1,1,1,1󰇞.
Eventually, to produce a true NURBS patch
which will continue to accurately represent the
circumference of the circle, an inner knot is inserted
at the middle of each knot vector in either direction.
Therefore, the twenty-five control points are
arranged in a 55 setup, as shown in Fig. 6(c), and
therefore the final knot vector becomes 𝛯"
󰇝0,0,0,0,
, 1,1,1,1󰇞 󰇝0,0,0,0,
, 1,1,1,1󰇞 while the
weights are 󰇝1, 0.9024, 0.8047, 0.9024, 1󰇞.
It is noted that the circle in green colour shown
in Fig. 6(b,c,d) is the accurate one which is ensured
in all the rational formulations of Model-1 and
Model-2.
The above-mentioned 25-DOF model (MODEL-
1, in a 55 setup, as shown in Fig. 6(c), is solved
first in conjunction with standard NURBS-based
IGA. This is accomplished once using the usual de
Boor functions (spcol in MATLAB) and another
using the equivalent Bézier extraction. Obviously,
the results were found to be identical, and the
associated errors are shown in the third column of
Table 1. It is noted that the aforementioned Bézier
extraction was based on increasing the multiplicity
of the nine inner knots (in a 33 setup) to 𝑝3,
thus the involved Bézier operator
e
C
implicitly
utilizes 49 control points.
Then, the 49-DOF model (MODEL-2, in a 77
setup), of which the control points had been
implicitly used in the aforementioned Bézier
extraction, is solved using four rational cubic Bézier
elements (separated by the two perpendicular thick
lines in red colour, as shown in Fig. 6(d)) and the
relevant results are shown in the fourth column of
Table 1.
Once again, one may observe that MODEL-2
outperforms in accuracy. Since the difference
between the degrees of freedom in the two Models
is less than 1:2, this problem is not the worst case.
Due to the above fact, in Section 3 the
comparison of CPU-times will be reported for only
the rectangular cavity as is explained therein.
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
611
Volume 17, 2022
Table 1. Eigenvalues of the circular cavity with hard
walls
Mode
EXACT
Errors (in %) of Calculated
Eigenvalues
MODEL-1
(25 DOFs)
MODEL-2
(49 DOFs)
1 0 - -
2 3.389957716671888 0.43 0.07
3 3.389957716671888 0.43 0.07
4 9.328363213746355 0.34 0.32
5 9.328363213746355 3.75 0.47
6 14.681970642123899 1.13 0.96
7 17.649988519749648 11.85 1.99
8 17.649988519749648 11.85 1.99
9 28.276371248725660 13.68 3.25
10 28.276371248725660 98.64 3.25
11 28.424282047372301 97.60 3.01
12 28.424282047372301 157.67 8.97
13 41.160133480153071 140.69 12.39
14 41.160133480153071 159.07 15.31
15 44.972222417793944 179.09 5.53
16 44.972222417793944 179.09 46.28
17 49.218456321694596 219.70 35.55
3 Discussion
If we restrict our discussion to the polynomial
degree 𝑝3, then MODEL-1 is of 𝐶-continuity
and leads to a pair of matrices, say (K1,M1),
whereas MODEL-2 is of 𝐶-continuity and leads to
another pair of matrices, say (K2,M2). Each pair of
matrices leads to a separate set of eigenvalues and
our findings suggest that the set (K2,M2) will lead to
lower (thus most accurate) values.
Actually, from the three numerical examples of
this study, it becomes clear that MODEL-1 (IGA of
𝐶-continuity) is less accurate than MODEL-2
(Bézier elements of 𝐶-continuity), when the
control points which are implicitly encountered in
the Bézier extraction (MODEL-1) are explicitly
used to form Bézier elements (MODEL-2).
One may advocate that the aforementioned fact
is due to the higher number of DOFs involved in
MODEL-2. Actually, in the 1D-problem the 6 DOF
of MODEL-1 increased (by 67 percent) to 10 in
MODEL-2. In the 2D-rectangular cavity the 40
DOF of MODEL-1 increased (by 180 percent) to
112, while for the circular cavity the 25 DOF of
MODEL-1 increased (by 96 percent) to 49 in
MODEL-2.
Of course, a fair comparison would require
extensive experimentation regarding the computer
effort and a relevant cost-benefit analysis, not only
for the Helmholtz equation but also for other partial
differential equations (PDEs) as well (e.g., Laplace-
Poisson, elasticity problems, electro-magnetics,
etc.). At the moment, for the examples of this study
which restricts to the Helmholtz equation, we have
found that MODEL-2 outperforms in accuracy but
its superiority in computer effort is sensitive and
depends also on the particular procedure followed
(see Procedures 2a2c, which were defined in sub-
section 2.1.2).
To become clear, we focus on the second
example (the rectangular cavity of subsection 2.2.1)
because it is characterized by the abovementioned
largest increase (180 percent) of the control points
between MODEL-1 and MODEL-2 (Procedure 2b)
thus it is the hardest case. To allow our work to be
reproduced by anyone interested, MODEL-1 (Bézier
extraction) was programmed in the MATLAB
environment cited in [29]. In the latter software only
slight modifications have been made in subroutine
FormK as follows: the Laplace operator () of our
study has substituted the Navier–Cauchy equations
of two-dimensional elasticity therein and thus has
led to a different stiffness matrix [K]. Moreover, a
mass matrix [M] has been added below the stiffness
matrix. The eigenvalues have been calculated using
the command eig(K,M). Furthermore, MODEL-2
has been programmed as an independent MATLAB
code in which Bernstein polynomials and their
derivatives were found using three alternatives
given in Table 2 under the header MODEL-2, as
follows:
1) Label spcol: The spcol function, which is
inherent in the Spline Toolbox of MATLAB,
was used to calculate the Bernstein
polynomials in each Bézier element,
independently.
2) Label Analytical: The four Bernstein
polynomials and their derivatives were
analytically calculated using the formulas:
𝐵, 󰇛1𝜉󰇜
, 𝐵
, 3𝜉
󰇛1𝜉
󰇜,𝐵
,
3𝜉󰇛1𝜉
󰇜,𝐵
, 𝜉
, etc.
3) Label Data: After the analytical computation
of Bernstein polynomials and their
derivatives, the numerical results were
tabulated in matrices of size 4×4 and were
put in the beginning of the subroutine which
estimates the element stiffness and mass
matrices.
In both models, the CPU-time has been
calculated as the mean average of 30 trials, on the
same computer under the same conditions, and the
results are given in Table 2. One may observe that
CPU-time required for the computer execution of
MODEL-2 is smaller than that required in MODEL-
1, even for the hardest (worst) case of using the
standard spcol function while it requires halve
computer time when the data at Gauss points are
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
612
Volume 17, 2022
given in advance. In any case, considering the more
accurate results of MODEL-2 shown in Fig. 5:, it is
concluded that MODEL-2 outperforms with respect
to MODEL-1 from all the points of view.
Table 2. CPU-time for a rectangular cavity
MODEL-1
MODEL-2
(Procedure 2a)
Bernstein polynomials using
s
p
col Anal
y
tical Data
0.088757 0.080333 0.055735 0.041925
The comparison in Table 2 is quite fair. In both
models the Jacobians have been calculated at the
4×4 Gauss points of all the ten B-spline or Bézier
elements shown in Fig. 4, accordingly. To make this
point still clearer, if –for instance– the element
matrices (𝐊,𝐌
) are analytically calculated in
MODEL-2 once for the first 16-node Bézier element
only, then the CPU-time dramatically reduces to
only 0.01328 seconds. Note that in MODEL-1 we
have to calculate 40 eigenvalues while in MODEL-2
we deal with 112 ones; so the CPU-time in
MODEL-2 includes extra computer effort for the
computation of a higher number of eigenvalues
(performed by the eig function).
While the abovementioned Procedure 2b (Bézier
element of 𝐶-continuity) clearly requires less
computer effort than the usual Bézier extraction, this
is not the case for Procedure 2d (Eqs. (6) and (7)).
Obviously, this is because Procedure 2d is a
continuation of MODEL-1. On this issue there is
still space for further contribution, with the hope
that the extra computer effort will be compensated
by the smaller numerical error. For example,
regarding the efficient computation of the two
involved quadratic forms


11
eeee
Q

KCKC
etc., one could apply ideas found in a recent study
on symmetric matrices, [30].
Ongoing research on programming details shows
that if we wish to calculate both models (i.e.
MODEL-1 and MODEL-2: Procedure 2a)
simultaneously, then at each Gauss point we have to
perform the Bézier extraction (MODEL-1) and then
calculate the Jacobian matrix in terms of the first set
of control points (see, [28]). But since after knot
insertion the shape of the domain remains
parametrically the same, the numerical values of the
aforementioned Jacobian matrix may be preserved
for MODEL-2 as well. The latter consists of using
the value of the Bernstein polynomials and their
local derivatives at the same Gauss points, a
procedure that can be done once outside the
subroutine which performs the numerical integration
of the matrices. Unlike the eigenvalue problem of
this study, regarding other types of PDEs in the
general form 𝐷󰇛𝑢󰇜 𝑓 0, the simultaneous
computation of the variable 𝑢 and its gradient ∇𝑢 at
the same points in both models (MODEL-1 and
MODEL-2: Procedure 2a) is promising to build a
new error-estimator for further adaptation of the
control points for a still more accurate numerical
solution.
Having already discussed the three Procedures
2(a, b and d), let us now turn to the remaining
Procedure 2c, which is related to tensor product
cubic Lagrange polynomials (here of uniform 4×4
nodes). This can be programmed as usual, following
standard FEM programming rules. The most critical
issue is that Lagrange polynomials and their local
partial derivatives should preferably be calculated in
advance at the Gauss points and stored, so the
computer effort substantially decreases.
Interestingly, Lagrange elements lead to identical
eigenvalues with Bézier elements of the same
degree. Here, we ought to explain the reason that
although Procedure 2b (Bézier elements) differs
from Procedure 2c (Lagrange elements) in their
matrices they both lead to the same numerical result
(not only in the eigenvalues of this paper but in all
other boundary-value problems). The explanation is
the fact that both functional sets of these
polynomials share the same monomials in the form
𝑥𝑦, so there is a linear relationship between
Bernstein-Bézier and Lagrange polynomials. In
other words, we only have a change of basis thus the
final numerical results become identical. On this
issue the interested reader may consult a detailed
discussion in [16, pp. 258-268].
In general, MODEL-1 should be carefully
programmed otherwise sometimes Bézier extraction
may even be slower than the usual IGA. But even
our programming is efficient the gain is not always
that big. For example, regarding a particular 2D
boundary value problem governed by Laplace
equation, and for control points which are tensor
product of Fig. 1 (i.e., 6×6 = 36 control points in
total), working with an older hardware we have
found that the required CPU-time for the usual IGA
is 0.971 sec, for the usual Bézier extraction is 0.928
sec, while for a smart implementation of Bézier
extraction is 0.900 seconds (7.3 percent reduction
with respect to the usual IGA). It is worthy to
mention that in the latter case the univariate
Bernstein-Bézier polynomials as well as their local
derivatives have been pre-calculated at the four
Gauss points in each direction once and then stored
for future demand (see, [31, p. 50]).
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
613
Volume 17, 2022
4 Conclusion
In the implementation of NURBS-based
isogeometric analysis (IGA), at this date Bézier
extraction is the standard procedure. The latter
requires that the multiplicity of inner knots increases
up to the polynomial degree, thus the Bézier
operator is calculated and then the basis functions
(of C2-continuity) are efficiently estimated. In this
paper it was shown that implementing trivial matrix
operations such as matrix inversion and
multiplications on the standard IGA stiffness and
mass matrices, it is possible to estimate larger
matrices which correspond to the involved Bézier
elements characterized by C0-continuity.
Interestingly, the same matrices may be derived
treating the Bézier elements independently, i.e., in a
similar way with the standard finite element method,
and this model showed to outperform.
APPENDIX A
Stiffness and mass matrices for the cubic Bézier
element of length e
l:
Using the well known formulas
0
e
l
ij i j
kBBdx

and 0
e
l
ij i j
mABBdx
, (Α-1)
in conjunction with the standard Bernstein
polynomials of degree three, i.e.,
32
12
23
34
(1 ) , 3(1 ) ,
3(1 ) , ,
BB
BB



 (Α-2)
it can be found that:
Bezier
Bezier
6321
34 1 2
3,
21 4 3
10
1236
60 30 12 3
30 36 27 12 .
12 27 36 30
420
3123060
e
e
l
l
















K
M
(Α-3)
APPENDIX B
Stiffness and mass matrices for the cubic
Lagrange element of length e
l:
Using the well known formulas
0
e
l
ij i j
kLLdx
and 0
e
l
ij i j
mALLdx
,
(Β-1)
in conjunction with the standard Lagrange
polynomials of degree three, i.e.,
9
1
22
91
22
12
34
(3 1)(3 2)(1 ), (3 2)( 1),
(3 1)(1 ), (3 1)(3 2)
LL
LL

 

 ,
(Β-2)
it can be found that:
Lagrange
Lagrange
148 189 54 13
189 432 297 54
1,
54 297 432 189
40
13 54 189 148
128 99 36 19
99 648 81 36 .
36 81 648 99
1680
19 36 99 128
e
e
l
l












K
M
(Β-3)
References:
[1] Taig I.C., Structural Analysis by the matrix
displacement method, Engl. Electric Aviation
Report No. S017, 1961.
[2] Irons B.M., Engineering applications of numerical
integration in stiffness matrix, J.A.I.A.A., Vol. 4,
No. 11, 1966, pp. 2035-2037.
[3] Zienkiewicz O.C., The Finite Element Method,
McGraw-Hill, 1977.
[4] Bathe K.-J., Finite Element Procedures, Prentice-
Hall, 1996.
[5] Gordon W.J., and Hall C.A., Transfinite element
methods: Blending-function interpolation over
arbitrary curved element domains, Numerische
Mathematik, Vol. 21, 1973, pp. 109–112.
[6] Cavendish J.C., Gordon W.J., and Hall C.A., Ritz-
Galerkin approximations in blending function
spaces, Numerische Mathematik, Vol. 26, 1976,
pp. 155–178.
[7] Farin G., Curves and surfaces for computer aided
geometric design: a practical guide, Academic
Press, 1990.
[8] Piegl L., andTiller W., The NURBS Book (1st ed.),
Springer, 1995.
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
614
Volume 17, 2022
[9] Sederberg T.W., Zheng J., Bakenov A., and Nasri
A., T-splines and T-NURCCSs. ACM
Transactions on Graphics, Vol. 22, No. 3, 2003,
pp. 477–484.
[10] Dem’yanovich Y.K., Belyakova O.V., and Le
B.T.V., General smoothness of the Hermite type
splines, WSEAS Transactions on Mathematics,
Vol, 17, 2018, pp. 359-368.
[11] Dem’yanovich Y.K., Miroshnichenko I.D., and
Musafarova E.F., On splines’ smoothness, WSEAS
Transactions on Mathematics, Vol, 18, 2019, pp.
129-136.
[12] Burova I.G., Application of non-polynomial
splines to solving differential equations, WSEAS
Transactions on Mathematics, Vol, 19, 2020, pp.
531-548.
[13] Garnier L., Druoton L., Bécar J.-P., Fuchs L, and
Morin G., Subdivisions of horned or spindle Dupin
cyclides using Bézier curves with mass points,
WSEAS Transactions on Mathematics, Vol, 20,
2021, pp. 756-776.
[14] Kanarachos A., and Deriziotis D., On the solution
of Laplace and wave propagation problems using
‘C-elements’, Finite Elemenets in Analysis and
Design, Vol. 5, 1989, pp. 97–109.
[15] Provatidis C., and Kanarachos A., Performance of
a macro-FEM approach using global interpolation
(Coons’) functions in axisymmetric potential
problems, Computers & Structures, Vol. 79, No.
19, 2001, pp. 1769–1779.
[16] Provatidis C.G., Precursors of Isogeometric
Analysis: Finite elements, Boundary elements and
Collocation methods, Springer, 2019.
[17] Kagan P., Fischer A., and Bar-Yoseph P.Z., New
B-spline finite element approach for geometrical
design and mechanical analysis, International
Journal for Numerical Methods in Engineering;
Vol. 41, No. 3, 1998, pp. 435–458.
[18] Kagan P., and Fischer A., Integrated mechanically
based CAE system using B-Spline finite elements,
Computer-Aided Design 2000; 32:539–552.
[19] Kagan P., Fischer A., Bar-Yoseph PZ.
Mechanically based models: Adaptive refinement
for B-spline finite element. International Journal
for Numerical Methods in Engineering 2003;
57:1145–1175 (DOI: 10.1002/nme.717)
[20] Höllig K., Finite element methods with B-splines,
SIAM, 2003.
[21] Schramm U., and Pilkey W.D., The coupling of
geometric descriptions and finite element using
NURBS—a study in shape optimization, Finite
Elem Anal Des, Vol. 15, 1993, pp. 11–34.
[22] Schramm U., and Pilkey W.W., Higher order
boundary elements for shape optimization using
rational B-splines, Eng Anal Boundary Elem, Vol.
14, No. 3, 1994, pp. 255–266.
[23] Inoue K., Kikuchi Y., and Masuyama. T., A
NURBS finite element method for product shape
design, J Eng Des, Vol. 16, No. 2, 2005, pp. 157–
174.
[24] Hughes T.J.R., Cottrell J.A., and Bazilevs Y.,
Isogeometric analysis: CAD, finite elements,
NURBS, exact geometry and mesh refinement,
Comput Methods Appl Mech Eng, Vol. 194, 2005,
pp. 4135–4195.
[25] Cottrell J.A., Hughes T.J.R., and Bazilevs Y.,
Isogeometric analysis: towards integration of CAD
and FEA, Wiley, 2009.
[26] De Boor C., On calculating with B-splines, J
Approx Theory, Vol. 6, 1972, pp. 50–62.
[27] Cox M.G., The numerical evaluation of B-splines,
J Inst Math Its Appl, Vol. 10, 1972, pp. 134–149.
[28] Borden M.J., Scott M.A., Evans J.A., and Hughes
T.J.R., Isogeometric finite element data structures
based on Bézier extraction of NURBS,
International Journal for Numerical Methods in
Engineering, Vol. 87, 2011, pp. 15–47.
[29] Nguyen T.N., Isogeometric Finite Element
Analysis based on Bézier Extraction of NURBS
and T-Splines, Master’s Thesis, Department of
Structural Engineering, Faculty of Engineering
Science and Technology, NTNU- Norwegian
University of Science and Technology, June 14,
2011. https://ntnuopen.ntnu.no/ntnu-
xmlui/handle/11250/236860
[30] Mitrouli M., Polychronou A., Roupa P., and Turek
O., Estimating the Quadratic Form 𝐱𝐀𝐱 for
Symmetric Matrices: Further Progress and
Numerical Computations, Mathematics, Vol. 9,
No. 12, 2021, 1432.
[31] Raditsas J., Performance of Bézier extraction in
isogeometric analysis, Master’s Thesis, National
Technical University of Athens, Post-graduate
course in “Applied Mathematical Scieces”,
February 2020 (Advisor: Prof. C. Provatidis, in
Greek).
https://dspace.lib.ntua.gr/xmlui/handle/123456789/
51215
Sources of Funding for Research Presented in a
Scientific Article or Scientific Article Itself
None.
Creative Commons Attribution License 4.0
(Attribution 4.0 International, CC BY 4.0)
This article is published under the terms of the
Creative Commons Attribution License 4.0
https://creativecommons.org/licenses/by/4.0/deed.en
_US
WSEAS TRANSACTIONS on SYSTEMS and CONTROL
DOI: 10.37394/23203.2022.17.67
Christopher G. Provatidis
E-ISSN: 2224-2856
615
Volume 17, 2022