You are on page 1of 13

International Journal of Crashworthiness

ISSN: 1358-8265 (Print) 1754-2111 (Online) Journal homepage: https://www.tandfonline.com/loi/tcrs20

Buckling considerations and cross-sectional


geometry development for topology optimised
body in white

J. Christensen , C. Bastien , M. V. Blundell & P. A. Batt

To cite this article: J. Christensen , C. Bastien , M. V. Blundell & P. A. Batt (2013) Buckling
considerations and cross-sectional geometry development for topology optimised body in white,
International Journal of Crashworthiness, 18:4, 319-330, DOI: 10.1080/13588265.2013.792442

To link to this article: https://doi.org/10.1080/13588265.2013.792442

Published online: 01 May 2013.

Submit your article to this journal

Article views: 199

View related articles

Citing articles: 3 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tcrs20
International Journal of Crashworthiness, 2013
Vol. 18, No. 4, 319–330, http://dx.doi.org/10.1080/13588265.2013.792442

Buckling considerations and cross-sectional geometry development for topology


optimised body in white
J. Christensen∗, C. Bastien, M.V. Blundell and P.A. Batt
Faculty of Engineering and Computing, Coventry University, Coventry, UK
(Received 11 January 2013; final version received 2 April 2013)

This paper will investigate how current state-of-the-art structural optimisation algorithms, with an emphasis on topology
optimisation, can be used to rapidly develop lightweight body in white (BIW) concept designs, based on a computer aided
design envelope. The optimisation models included in the paper will primarily focus on crashworthiness and roof crush
scenarios as specified in the Federal Motor Vehicle Safety Standards (FMVSS) 216 standard. This paper is a continuation
of a previously published paper, which investigated the potential effects of recently proposed changes to FMVSS 216 upon
BIW mass and architecture using topology optimisation. The paper will investigate the possibilities of including buckling
considerations of roof members directly into current state-of-the-art topology optimisation algorithms. This paper will also
demonstrate the potential for developing a detailed BIW design including cross-sectional properties based on a styling
envelope.
Keywords: finite element topology optimisation; FMVSS 216; roof crush; body in white (BIW); lightweight vehicle
architecture; buckling

1. Introduction cases based on linear static finite element (FE) modelling.


In 2003, it was estimated that approximately 10,000 fatali- As the magnitudes of the applied forces did remain constant
ties occurred on US roads due to rollovers [12]. This high- throughout the study [5], the effects of the change of angles
lights the severity of the crashes associated with rollovers; could be accommodated by changing the load paths, with
the influence of the roof strength is discussed in [10]. In light only minor changes to overall BIW mass.
of this, recent changes to the FMVSS 216 standard has been The effects and importance of buckling and localised
proposed. The potential effects of these changes upon body crushing of the roof pillars as significant factors in the
in white (BIW) architecture and mass have previously been overall crash performance of the roof have been discussed
investigated using topology optimisation, inertia relief and in [4], thereby making these potential parameters of topol-
a hybrid electrical vehicle case study [5]. The only variation ogy optimisation. Using commercially available topology
between the individual models of the study in [5] was the optimisation software, it is not feasible to directly imple-
(pitch and roll) angles associated with the roof crush sce- ment the above parameters as variables into optimisation
nario. The pitch angle  and the roll angle θ are illustrated models.
in Figure 1. Furthermore, the applied loading magnitude Firstly, buckling cannot be calculated for solid (three-
was increased to three times the unloaded vehicle weight dimensional (3D)) elements. Secondly, it is important to
(45 kN), i.e. excluding fuel, passengers etc., as suggested recognise the ‘relative simplicity’ of the models in ques-
in [3,11,13,14]. tion. In general, the outcome of FE-based topology optimi-
The study in [5] included 11 models and found that the sation does not contain detailed information relating to e.g.
estimated mass value of the roof, denoted as 1 in Figure 2, cross-sectional geometry, which is required in order to draw
did not vary significantly as a function of the pitch and roll accurate estimations on the likelihood of buckling and/or
angle variations. localised crushing of the roof pillars occurring.
This may initially seem controversial, as other pa- To include buckling and localised crushing (with a sat-
pers investigating the effects of the proposed FMVSS 216 isfactorily level of accuracy), as parameters in connection
changes, such as [4], found that the magnitude of force dis- with topology optimisation, alternative optimisation algo-
tribution within the roof changes significantly as a function rithms, such as the homogenisation method, could possibly
of the pitch and roll angles. However, at this point, it is be employed [2]. However, these algorithms have proven
important to remember that topology optimisation extracts extremely difficult, if not impossible to apply to anything
the most ‘efficient’ load paths according to the applied load but very simple geometries, let alone a ‘complex’ design


Corresponding author. Email: aa8867@coventry.ac.uk


C 2013 Taylor & Francis
320 J. Christensen et al.

Figure 1. Pitch and roll angle according to FMVSS 216.


Figure 3. Example of roof topology; relative element density.
volume, such as the one illustrated in Figure 2 (including a
large number of elements and design space based on com-
plex Boolean operations). these beams, due to packaging space, i.e. design envelope,
The implementation of buckling and localised crush- the only remaining option is to increase the material gauge,
ing as parameters in topology optimisation is not certain to inadvertently leading to an increase in mass, disregarding a
lead to significant changes in the results of the FMVSS 216 possible change of material.
study of [5]. This is primarily linked to the objective of the On the other hand, if the aspects and limitations stated
optimisation, which was to minimise the BIW mass whilst above were considered and actively used during the topol-
constrained by maximum displacement criteria. It there- ogy optimisation process, the general outcome (topology)
fore follows that the optimisation will define/retain the load may be significantly different. This is based on the fact
paths where they are most ‘efficient’. Consequently, the that ‘shorter beams’ will be less likely to experience buck-
optimisation will attempt to maximise the forces in the in- ling, hence I may be decreased. This is, however, likely to
dividual load paths, inadvertently increasing the likelihood considerably increase the complexity of the roof structure,
of buckling. possibly invoking manufacturing constraints. Finally, it is
This fact raises an interesting question with respect far from certain that the overall mass of the ‘short beam’
to obtaining the lightest roof structure. To prevent buck- structure is less than that of the ‘long beam’ structure, this
ling of ‘long beams’, i.e. the roof members, obtained using warrants further analysis and is the main focus of atten-
commercial software, as illustrated in Figure 3 (displaying tion of this paper. The problem will be approached using
relative element density), the cross sections of these will existing topology optimisation algorithms as used in [5].
subsequently have to be dimensioned accordingly during Figure 4 illustrates the proposed overall algorithm for
the optimisation process. BIW development using structural optimisation. The main
This does in essence mean that (for a constant value of focus of this paper will be step 3 of Figure 4. Before this is
Young’s Modulus) the second moment of area I, needs to be completed step 2 as was originally completed in [5] will be
defined. As there is a clear limit to the outer dimension of revisited; I value estimations. There are some immediate

Figure 2. Roof area definition.


International Journal of Crashworthiness 321

Table 1. Selected models for case study.

Model # Pitch angle (◦ ) Roll angle θ (◦ )

1 0 0
4 10 40
6 5 45
7 10 45

maximum mass difference between all models of this study


was found to be less than 0.5% [5]. It is also worth not-
ing that other studies, such as [1,3,13], have found that the
Figure 4. Overall proposed optimisation methodology. combination of a 10◦ pitch angle , and a 45◦ roll angle θ
constitutes the ‘worst case loading scenario’.
complications of how the I-value estimation could be made To evaluate the likelihood of buckling eigenvalue based
an active parameter within the topology optimisation algo- algorithms are commonly used in FE. However, due to the
rithm, let alone as a constraint. This will be the focus of usage of inertia relief in connection with the topology op-
attention in the following section. timisation, it is not feasible to use an eigenvalue-based ap-
proach as the two methods are incompatible. Therefore, it
was chosen to use Euler’s buckling formula to estimate the
critical buckling load, F crit , Equation (1),
2. Implementation of I-value
π2 · E · I
To fully implement an accurate I-value within topology Fcrit = , (1)
optimisation a meta-modelling technique similar to that k · L2
used in the shape and size optimisation of [9] could be where E is Young’s modulus, I is the second moment of area,
adopted. However, this approach was deemed infeasible k is the slenderness ratio and L is the length of the beam
in this connection, particularly as topology optimisation member. For the remainder of this paper, k will be set to 1.
determines the general load paths without providing high The length L of each beam member was predetermined (and
levelled details of cross-sectional properties. fixed) based on the outcome of the topology optimisation
Therefore, it was much more feasible to include an results. L was determined on a member basis (based on
approximation of the local I-value within the topology op- a constant mesh size), as opposed to an element basis, as
timisation algorithm. To do so, the following points of con- illustrated in Figure 5.
sideration needed to be addressed: For the purpose of conducting the size optimisation,
in order to estimate the I-values, a tubular cross section
(1) Establish how the local I-values could be estimated. of unity thickness was selected for convenience, as this
(2) Determine how the I-values could be used to evaluate left only one variable parameter: the diameter d. It has to
the probability of the occurrence of buckling. be noted that the general cross-sectional geometry (shape)
(3) Define how to implement step 2 into the topology opti- was irrelevant, as the impending task was to estimate ap-
misation algorithm. propriate I-values for buckling estimations. With the above
definitions, the buckling factor, Bf , could be estimated using
To determine a feasible methodology for estimating the
appropriate I-values for the ‘local roof beams’, a case study
was performed. The intention was to use size optimisation
of a fixed cross-sectional geometry, i.e. no shape optimi-
sation, and include the buckling factor as a constraint. To
do so, wireframe models of representative topologies of
four models from the results of the topology optimisation
study [5] were used. The four selected models are listed in
Table 1.
It is important to note that the selected models in
Table 1 contain the current EuroNCAP set-up ( =
θ = 0◦ ); this grouping also included the current
FMVSS 216 set-up [5]. Model 7, also listed in
Table 1, represented the highest mass reduction of the orig-
inal FMVSS 216 topology optimisation study, although the Figure 5. Example of wireframe model.
322 J. Christensen et al.

Figure 6. Results from diameter sizing.

Equation (2), roof sheets, as illustrated in Figure 6. The values are not to
be interpreted as the final (ready for production) masses, as
Element axial force shape-optimisation has not been completed as this stage. In
Bf = < 1. (2) fact another sizing optimisation will need to be completed
Fcrit
at a later stage in order to further optimise the structure. It is
The ‘element axial force’, used in Equation (2), is deter- important to remember that the overall purpose of the above
mined by the maximum axial force occurring within any was to investigate the first step of the potential for including
element of the member in question. As long as Bf re- buckling considerations within topology optimisation, i.e.
mains less than 1, linear buckling is not likely to occur. step 2 of Figure 4.
The above methodology was relatively simple to include as Furthermore, the wireframe models included in the
a constraint within a sizing optimisation, using commercial study (Tables 1 and 2) were kept as close to the ‘original’ as
software. possible, in particular this meant the inclusion of a signifi-
The results from the four models listed in Table 1, are cant number of ‘curved beams’, as indicated by Figure 5 and
listed in Table 2, and illustrated in Figure 6. Figure 6. Revising the topology prior to ‘I-value determi-
The estimated masses listed in Table 2 include A-, B-, C- nation’ may have a significant effect on the estimated mass
and D-pillars as well as the main roof structure, excluding results in Table 2. Nevertheless, the mass results listed in
Table 2 may be compared relative to each other in order to
better understand the ‘relative influence’ of the pitch and
Table 2. Results from I-value estimation (sizing optimisation). roll angles upon the roof structure. On this basis, it can
be seen that the current EuroNCAP test,  = θ = 0◦ , is
Pitch Roll Final of ‘medium severity’, based on the final mass estimation.
angle angle estimated
Model # (◦ ) θ(◦ ) mass (kg) Even more interesting model 7 ( = 10◦ , θ = 45◦ ) rep-
resents a significantly higher estimated mass value with
1 0 0 42.1 a difference of approximately 11% relative to model 6
4 10 40 41.0 ( = 5◦ , θ = 45◦ ) highlighting the necessity for subse-
6 5 45 39.7 quent sensitivity studies as discussed in [5]. This is in
7 10 45 44.4
line with the findings of [1,3,13], who established that the
International Journal of Crashworthiness 323

‘worst crash loading scenario’ occurs when  = 10◦ and


θ = 45◦ , thus substantiating the appropriateness of the re-
sults obtained by means of topology optimisation and the
subsequent I-value estimation by scaling the diameters.
This does, however, not solve the previously proposed
problem of a ‘long beam structure’ versus a ‘short beam
structure’, as the I-value is only considered subsequent to
the completion of the topology optimisation. The next steps
of how this could be achieved will be the focus of attention
in the following section.

2.1. Implementation of I-value estimation within


topology optimisation
Although the results from the size optimisation estimat-
ing the local I-values suggestively produce reasonable
Figure 7. Example of topology iteration; relative element density.
approximations they cannot straightforwardly be imple-
mented within a topology optimisation algorithm, such as
the one employed in [5]. Using a mathematically based
rent (iteration) values of the relative element densities. If
topology optimisation algorithm and the solid isotropic ma-
this problem was to be solved, it would also aid in the of-
terial with penalisation (SIMP) interpolation scheme two
ten lengthy task of post-processing the optimisation results,
primary problems needed to be addressed; the identifica-
by offering a certain level of ‘automation’ with the aim
tion of ‘beams’, i.e. load paths, and how to include the
of obtaining a computer-aided design (CAD) model to be
I-value estimations as active parameters within the algo-
used for further analysis, or indeed additional structural op-
rithm. At this point, it is important to highlight that it was
timisation. Implementation of this feature would, however,
necessary to use static modelling as 3D (solid) elements
require disassembly of the commercial FE solver, which is
were used for the topology optimisation [5]. Therefore, it
prohibited and therefore not possible at this stage. Follow-
was not feasible to consider using one or several (explicit)
ing the identification of load paths the next step would be to
dynamic crash analyses. This is due to the fact that the
include the I-value estimations as active constraints within
central processing unit time would dramatically increase
the topology optimisation.
as a function of the small timestep (t); a consequence
of the small element size required for the 3D elements.
To increase t element mass could be added; this would, 2.2. Active estimation of I-values
however, need to be of significant magnitude making the FE Assuming that the above problem was solved, the next step
model incompatible with the real world physics of the crash would then be to implement the I-value estimations within
event. the topology optimisation algorithm. This would initially
Consequently linear static (implicit) finite element anal- not pose a significantly challenging problem, as this could
ysis (FEA) would be the starting point for the I-value esti- simply be obtained by means of a constraint imposed within
mations. the topology optimisation problem definition, in the exact
With the starting point defined, the next part-task for the same manner as e.g. a displacement constraint. In this case,
I-value estimations could be approached. This was the issue the methodology proposed in the previous section could be
of how to ‘define/identify’ the load paths within a specific straightforwardly employed.
iteration of the topology optimisation process, in order to However, this would theoretically neither provide a sig-
model these as ‘beams’ for estimation of appropriate I- nificant engineering nor scientific advancement as the end
values. This was, however, a very complex task, which results would most likely not differ considerably from those
may be envisaged by observing the example in Figure 7, illustrated in Figure 6 or those listed in Table 2. The rea-
illustrating relative element density. son for this is simply that such an approach would merely
The example in Figure 7 clearly demonstrates the com- include the local I-values as ‘inactive constraints’, and not
plexity of (explicitly and uniquely) identifying the primary actively guiding/influencing the topology optimisation pro-
structural load paths within any iteration. Considering cur- cess or results.
rent state-of-the-art modelling/optimisation algorithms, it If the I-value estimations were to be actively included,
is believed that this proposed technique has the highest the task at hand would become significantly more com-
potential to resolve this issue by means of logical math- plex, as the ‘fundamental’ decision-making process of the
ematical expressions embedded within the FE solver, i.e. topology optimisation algorithm would have to be rede-
observing and using Boolean expressions based on the cur- fined [1,2]. The fact that the I-value estimations include the
324 J. Christensen et al.

Figure 8. Representation of FE model.

length of the beam, which would not be defined at that stage 2.3. Substantiation of results
of the topology optimisation process, adds further complex- Prior to generating (manufacturable) local cross sections,
ity to the problem, as previously discussed and illustrated the results obtained via the topology and subsequent size
by Figure 7. optimisation were substantiated via dynamic (explicit) non-
The two aspects discussed above thus make it ex- linear FE modelling. This was conducted using the informa-
tremely difficult to implement the I-value estimations, and tion from the topology optimisation as well as the I-value
indirectly the buckling considerations, as an active param- estimations, to create (two-dimensional (2D)) shell models,
eter within a topology optimisation algorithm. Evaluating as illustrated in Figure 6.
whether or not solving the extremely complex tasks dis- The four models were separately meshed, using an ele-
cussed above would be worthwhile taking on was in essence ment size of 4 mm, leading to the creation of approximately
a question of estimating the potential effects of ‘long fat 360,000 elements per model. The models used the LS-Dyna
beams’ versus ‘short thin beams’ as previously discussed. ‘MAT 003: Plastic Kinematic’ material card with the pa-
In this context, it was deemed that the influence upon the rameters listed in Table 3.
final roof design would most likely be negligible, partially A rigid plate was subsequently inserted above the
due to manufacturing limitations in terms of cost versus the A-pillars in all four models, with the angles defined in
potential mass reduction in percent. In addition, it needed Table 1. During the analysis, these rigid plates were sub-
to be considered that the topology optimisation was com- jected to a prescribed displacement of 250 mm perpendic-
pleted using linear static FEA, as discussed in [5–9,15]. ular to the surface of the plate, i.e. moving in the yz-plane,
Furthermore, the Eulerian buckling formula, Equation (1), Figure 8, with a velocity of 5 mph [3], thus simulating a roof
assumes linear material behaviour. Thereby the method of crush scenario. The model set-up is illustrated in Figure 8.
superposition was in principle applicable, as the penalisa- The contact force between the rigid plate and the roof
tion factor p equalled 1 in [5]. Consequently it was unlikely structure was measured during the analyses. This could
that any significant difference in results would be found subsequently be compared to the applied force of 45 kN
between completing the I-value estimations subsequent to used for the topology and size optimisations [5], in order
the topology optimisation, or including these directly in the to evaluate the performance of the roof structure relative
topology optimisation step, provided that the latter were not
actively guiding the optimisation process, further analyses
are, however, required to support this claim. It therefore be- Table 3. Material parameters for explicit analysis.
came clear that if the buckling phenomena were to be ‘fully Parameter Value SI unit
implemented’ into a topology optimisation algorithm, using
3D elements, the fundamental principles of the algorithm Young’s modulus (E) 210,000 MPa
would have to be revised, as well as disassembly of the FE Poisson’s ratio (ν) 0.3
kg
solver. Volumetric mass 7850
In light of the above, it was deemed acceptable to com- density (ρ) m3
plete the I-value estimations subsequent to the topology Plastic tangent 1000
hardening modulus
optimisation as previously completed and listed in Table 2
(ETAN)
and illustrated by Figure 6.
International Journal of Crashworthiness 325

Figure 9. Results of explicit FE analysis, force vs. time.

to the original requirements. Figure 9 displays the contact be reiterated that the structure has been derived based on
force (between the rigid plate and the roof structure) as a topology optimisation incorporating; front, rear, side and
function of time. Figure 9 substantiates the validity of the pole crash scenarios, which inevitably have influenced the
approach, as the peak contact force of all four individual topology of the structure. The results of the FE analyses
models exceed 45 kN. thus substantiate that the obtained structures perform as
Figure 10 illustrates the deformed roof (model 7) at anticipated, the next step is therefore to generate (detailed)
t = 0.15 s, i.e. at a rigid plate displacement of 250 mm. The local cross-sectional geometry.
results in Figure 10 indicate that the roof structure does not
exhibit significant buckling. The figure also indicates that 3. Generation of local cross-sectional geometry
only localised deformation occurs, indirectly suggesting To generate the detailed local cross sections of the roof, in
that the structure is ‘over dimensioned’. It must, however, preparation for manufacturing, an optimisation algorithm

Figure 10. Deformed roof structure, model 7, t = 0.15 s.


326 J. Christensen et al.

Figure 11. Cross section of metal sheet.

Figure 12. Segments for cross section.


was created. This did also build on the outcome of the
topology optimisation as well as the I-value estimations.
The first step in this process was to define a roof segment 3.2. Six segment model
model to allow for the cross sections of the roof to vary in The cross section of a modern day vehicle does not only
size, and shape, throughout the structure, with the aim of consist of a single segment. Typically six segments are used
minimising the roof mass. as illustrated in Figure 12. These are defined by the letters
A–G, excluding C which denotes the centroid of the entire
cross section. The local (segment i) coordinate axes are
3.1. Segment model denoted Yi and Zi , αi denotes the angle of rotation between
As roof members are typically manufactured from (metal) the local and global coordinate system. It was assumed that
sheets, it was decided that ‘conventional’ manufacturing t remains constant for all segments; as previously stated the
processes would be used in connection with the roof de- roof members were to be manufactured using metal sheets
sign. The cross section of each roof member would consist of uniform thickness.
of a series of segments. The cross section of a flat metal The coordinates for the centroid C were obtained by
sheet would consist of one single segment, as illustrated in Equation (4),
Figure 11.
The segment example in Figure 11 is located in the
1 
6
yz-plane, is defined by the two endpoints A and B with YC = ASi · YCi
the coordinates (YA , ZA ) and (YB , ZB ), respectively. The ACS i=1
(4)
coordinates of the centroid C of the segment are (YC , ZC ). 1 
6
Finally, the thickness of the segment is denoted t and the ZC = ASi · ZCi ,
ACS i=1
angle between the segment and the y-axis is denoted α. The
area of the segment, the angle α and the coordinates for
the centroid C can all be obtained providing A, B and t are where ACS is the area of the entire cross sections, ASi is the
known parameters. area of the individual segment, Figure 11, whilst YCi and
Using the parallel axis theorem, the expressions for the ZCi are the coordinates of the centroid for the individual
second moment of area around the centroid C with respect segment. The second moments of area for the entire cross
to the global coordinate system axes can be defined as section can thus be obtained by firstly determining these for
expression (3), the individual segment i, with respect to the global coor-
dinate system using the parallel axes theorem, expression

ICy1 + ICz1 ICy1 − ICz1  (5),
ICY = + cos (2α) 
2 2  ICy1
ICy1 + ICz1 − ICz1
cos (2α) 
ICy1
ICZ = + (ICY )i = (ICi Y )i + SAi · (dZ )2
2 2 , (5)
(ICZ )i = (ICi Z )i + SAi · (dY )2
t · lAB
3 (lAB ) · t
3
= ∧ ICy1 = , (3)
12 12 where dY and dZ are the relative distances from the cen-
troid of the individual segment to the centroid of the com-
where lAB is the length of the line segment between point A bined cross section. The second moments of area for the
and point B. entire cross section can subsequently be obtained using
International Journal of Crashworthiness 327

Figure 13. Initial design variables for shape optimisation.

expression (6), points in order to then extract the minimum section mass.
Indeed, computing a response surface from more than 1000

6 points was not an efficient method of optimisation, as it
ICY = (ICY )i used a significant amount of disk space and caused an ex-
i=1
. (6)
6 cessive memory shortfall when DOE response points had
ICZ = (ICZ )i to be reloaded for the computation of the response surface.
i=1
The direct optimisation method was mostly chosen because
Thereby the second moment of area for any cross section solving each optimisation in an Excel spreadsheet took just
can be calculated regardless of the number of segments a few seconds with acceptable results, as will be discussed
within the cross section. later.
The methodology presented above was subsequently To conduct the optimisation, it was necessary to define
programmed into a spreadsheet, which could be used to the design variables including initial values and constraint
immediately calculate the second moments of area for any magnitudes. The majority of these would have to be set on a
cross section. The calculations of the spreadsheet were sub- local level, i.e. for each roof member (section); however, pa-
sequently validated using commercial CAD and FE soft- rameters such as the mass of the roof would be monitored
ware as well as manual calculations. Following this, the on a global level. Figure 13 represents a ‘typical’ cross-
spreadsheet was set up to be able to calculate and store the sectional roof profile and can be used as a starting point for
second moments of area for a series of cross sections in con- an example to define design variables. The coordinates of
nection with each other, which could be used to calculate points A–G (excluding C) were potential design variables
these for the entire roof structure. for the optimisation process. In this particular case, A, B
and D were fixed as these in essence define the external
profile, i.e. the A-surface of the vehicle. Thereby three de-
3.3. Optimisation of cross sections sign variables remained, namely the coordinates of points
E, F and G. There were, however, some restrictions on how
With the ability to calculate the second moments of area for
these would be able to change, which was primarily related
all roof members, these could be equated to the outcome
to the resulting interior cabin space (head room).
of the individual I-value estimations discussed in Section
1, illustrated in Figure 6. It should also be noted that parts
of the cross section will follow the outer profile, i.e. the A-
surface of the vehicle. This meant that the I-value could be
used as a constraint in connection with shape and size op-
timisation with the objective of minimising the roof mass.
Alternatively, the objective could also be to maximise the
I-values of the cross-sections, as the anticipated change in
roof mass was low. Nevertheless, the ultimate aim of the
exercise was to define manufacturable cross-sections of the
roof. The optimisation was to be conducted using commer-
cial software, which aimed to use design of experiments
(DOE), as described in [9] in order to build the response
surface of the vehicle cross section. Due to the vast number
of design variables and design variable levels, it was found
advantageous to perform a direct optimisation eliminating
the need to compute the response surface from experimental Figure 14. Updated design variables for shape optimisation.
328 J. Christensen et al.

Table 4. Parameters for initial shape and size optimisation.

Design variables YE , ZE , YF , ZF , YG , ZG, R1 , R2 , R3,


TR1 , TR2 , TR3 , TR4
R3, TR1 , TR2 , TR3 , TR4
Optimisation responses ACS , ICY , ICZ
Optimisation constraints ICY = fixed; ICZ = fixed
Optimisation objective Minimise ACS

blend surfaces would have to be used. The optimisation


parameters are summarised in Table 4.
Based on the information summarised by Table 4, the
optimisation problem was set up using Microsoft Excel
coupled with Altair HyperStudy to directly solve the op-
Figure 15. Optimised cross section. timisation problem. Using the roof rail as a case study, it
was found that the most efficient solution, given the de-
In addition to the actual coordinates of E, F and G, sign variable in Table 4 used the reinforcement R1 with a
the gauge thickness of the individual segments (1 to 6) gauge thickness (TR1 ) equal to 1.0 mm; this scenario met all
could also be design variables. However, as modern day (mathematical) optimisation constraints, including match-
roof structures usually use a minimum steel gauge thickness ing the I-value to that of the previous estimation. Despite
of 0.7 mm for reasons other than crash performance, it was this, there were concerns about the manufacturability of
decided to set the gauge thickness of the six segments to the proposed design; hence, it was decided to insert two
0.7 mm. additional points (H and J ) to guide the profile of the rein-
Figure 13 also displays four potential reinforcements forcement R1 , the updated design variables are illustrated in
(R1 , R2 , R3 and R4 ) which could be ‘inserted’ or Figure 14. With these changes the optimisation process was
‘removed’ using simple Boolean operators. Furthermore, repeated and a manufacturable design was obtained. This
the gauge thickness of the four potential reinforcements, consisted of three parts: the inner panel, the outer panel and
TR1 , TR2 , TR3 and TR4 were also included as design vari- the reinforcement, as illustrated by Figure 15.
ables, which could either assume the discrete values of To substantiate the feasibility of the proposed man-
1.0 mm or 0.7 mm. ufacturing methodology with respect to the design in
Finally manufacturing constraints were taken into ac- Figure 15, the geometry was analysed using FE-based sheet
count during the optimisation process; it was assumed that metal forming analysis (Altair HyperForm). This analysis
the roof structure would be tack welded and use a sheet used 2D (shell) elements with an average mesh size of
metal forming process, which meant that blank holders and 5 mm; the material parameters used are defined in Table 5.

Figure 16. Results from sheet metal forming analysis of optimised cross section.
International Journal of Crashworthiness 329

Table 5. Material parameters for sheet forming analysis. that the buckling calculations could be performed subse-
quently to the completion of the topology optimisation.
Parameter Value SI unit
Consequently, an ‘I-value’ estimation study was set up
Young’s modulus (E) 210,000 MPa using the load paths from the topology optimisation and
Poisson’s ratio (ν)   0.3 circular tubes with a gauge thickness of 1.0 mm to esti-
Yield stress value σy 315 MPa mate the required I-values of each roof member in order
Friction coefficient (FC) 0.125 to successfully withstand buckling failure in the event of a
vehicle rollover.
Dynamic (explicit) non-linear FE analysis was used to
Figure 16 displays the relative thinning of the sheets after substantiate the structural performance of the obtained load
the forming process. path and I-values. The results of these analyses clearly
Figure 16 displays relatively minor variation in thick- indicated that the structures obtained were fully capable of
ness for all three parts analysed. withstanding the applied loading.
The part found to have the largest variation of gauge Subsequently, a combined shape and size optimisation
thickness, as a function of the forming process was the outer study was set up in order to derive manufacturable cross
panel. In this case, the maximum increase (of thickness) was sections based on the topology optimisation and I-value es-
0.429% and the minimum decrease was −0.235%, mean- timations. Finally, the manufacturability of these cross sec-
ing a maximum thickness of 0.703 mm and a minimum tions were substantiated using FE-based sheet metal form-
thickness of 0.698 mm, based on an initial gauge thickness ing analysis to estimate the variations of gauge thickness,
of 0.7 mm. All elements of the analyses were below the and indirectly the structural performance of the manufac-
failure level of the forming limit diagram. tured parts.
The results of the sheet forming analyses thus sub- In conclusion, this paper has demonstrated the potential
stantiate that the parts can be manufactured without any for using current state-of-the-art commercially available
significant variations of thickness, indicating that the antic- optimisation software to obtain a manufacturable BIW able
ipated structural performance of the previously conducted to meet crash performance requirements, based on a styling
FE analyses can indeed be obtained. (CAD) surface.
The paper also recognises that revised (topology) opti-
4. Conclusion and next steps misation algorithms would be very beneficial in the effec-
The starting point for this paper was complications un- tiveness and validity of the overall methodology.
covered during a previous study, which used topology op-
timisation to investigate the potential effects of proposed
References
changes to FMVSS 216 upon BIW architecture [5]. The
[1] Altair Engineering, Inc. Altair Hyperworks Manual, Altair
use of topology optimisation enables the possibility of sig- Engineering, Inc., Troy, MI, 2010.
nificant reductions in BIW mass, particularly in connec- [2] M.P. Bendsoe and O. Sigmund, Topology Optimization, The-
tion with ‘alternatively’ fuelled vehicles, such as HEV. Al- ory Methods and Applications, Berlin, 2004.
though the outcome of the study suggested that currently [3] E.C. Chirwa, and Q. Peng, Modelling of roof crush using
available topology optimisation software can be used for the newly updated FMVSS 216, ICRASH, Washington, DC,
2010.
BIW load path extraction, it was recognised that buck- [4] E.C. Chirwa, R.R. Stephenson, S.A. Batzer, and R.H. Grze-
ling effects, which is the primary failure mode for rollover bieta, Review of the Jordan rollover system (JRS) vis-à-vis
events, were not taken into account. The study in this pa- other dynamic crash test devices, Int. J. Crashworthiness 15
per has investigated various options for considering buck- (5) (2010), pp. 553–569.
ling as part of the optimisation process. Initially, it was [5] J. Christensen, C. Bastien, and M.V. Blundell, Effects of roof
crush loading scenario upon body in white using topol-
investigated how this could be included (as an active pa- ogy optimisation, Int. J. Crashworthiness 12 (1) (2012),
rameter) during the topology optimisation process. This pp. 29–38.
approach was, however, deemed unfeasible using current [6] J. Christensen, C. Bastien, M.V. Blundell, A. Gittens, and
commercially available (topology) optimisation algorithms, M. Dickison, Integration of electric motor and alternator in
as the detailed information related to the geometry (of smart lightweight vehicles, Proceedings of the 4th Interna-
tional Conference on Mechanical Engineering and Mechan-
the roof sections) are not available during the topology ics, 2011, pp. 921–932. ISBN978-1-933100-40-1.
optimisation. [7] J. Christensen, C. Bastien, M.V. Blundell, A. Gittens, and
As buckling could not be included as an ‘active param- O. Tomlin, Lightweight hybrid electrical vehicle structural
eter’ within topology optimisation, the potential for im- topology optimisation investigation focusing on crashwor-
plementing it as a ‘passive’ parameter was investigated. thiness, Int. J. Vehicle Struct. Syst. 3 (2) (2011), pp. 113–122.
[8] J. Christensen, C. Bastien, M.V. Blundell, O. Grimes, A.
Recognising that topology optimisation uses linear static Apella, G. Bareham, and K. O’Sullivan, Modelling of
FEA (and a penalisation factor of 1 was used in [5]), the lightweight hybrid electric vehicle architectures in case of
method of superposition was deemed applicable, meaning the newly updated FMVSS 216 roof crush scenarios, The
330 J. Christensen et al.

International Crashworthiness Conference, ICRASH, Mi- cle rollover propensity, Int. J. Crashworthiness 8 (1) (2003),
lano, Italy, 2012. pp. 63–72.
[9] J. Christensen, C. Bastien, M.V. Blundell, and J. Kurakins, [13] R.R. Stephenson, The case for a dynamic rollover test,
Lightweight body in white design using topology-, shape and ICRASH, Washington, DC, 2010.
size optimisation, Electric Vehicle Symposium (EVS26), [14] F. Tahan, K. Digges, and P. Mohan, Sensitivity study of
Los Angeles, California, 2012. vehicle rollovers to various initial conditions – finite el-
[10] D. Friedman, and C.E. Nash, Measuring rollover roof ement model based analysis, ICRASH, Washington, DC,
strength for occupant protection, Int. J. Crashworthiness 8 2010.
(1) (2003), pp. 97–105. [15] J. Trollope, C. Bastien, J. Christensen, C. Kingdom, and M.V.
[11] R.H. Grzebieta, A.S. McIntosh, and M. Bambach, How Blundell, Optimisation of a common vehicle architecture in-
stronger roofs prevent diving injuries in rollover crashes, cluding electric vehicle, hydrid electric vehicle and inter-
ICRASH, Washington, DC, 2010. nal combustion engines propulsion systems for durability
[12] S.A. Richardson, G. Rechnitzer, R.H. Grzebieta, and E. and crash performance, The International Crashworthiness
Hoareau, An advanced methodology for estimating vehi- Conference (ICRASH), Milano, Italy, 2012.

You might also like