You are on page 1of 146

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/349427879

Regulated safesty structures of a formula student composite monocoque

Thesis · August 2020


DOI: 10.13140/RG.2.2.17138.63683

CITATIONS READS

0 1,392

1 author:

Sebastian Maklary Rudbeck-Rønne


Aquila Racing Cars A/S
2 PUBLICATIONS 0 CITATIONS

SEE PROFILE

All content following this page was uploaded by Sebastian Maklary Rudbeck-Rønne on 19 February 2021.

The user has requested enhancement of the downloaded file.


7

Faculty of Engineering
University of Southern Denmark

Regulated Safety Structures of a


Formula Student Composite
Monocoque
Master thesis

Author:
Sebastian Maklary Rudbeck-Rønne Supervisors:
(formerly Jørgensen) Assistant. Prof. Andreas
Project period: Mehl

Feb 2019 to Aug 2020 Associate. Prof. Vikas


Arora
Extent: 40 ECTS points
Version 2.3
90,33 pages of 2400
characters at hand in. August 20th, 2020
Preface

This thesis is part of the master’s degree Structural Engineering at the University of Southern
Denmark and has been written by Sebastian Maklary Jørgensen. During the span of this project, the
author has become a father and the COVID-19 crisis occurred, which extended the project deadline
from the initial date of January 2020 to August 2020. As this extension is due to the events
prohibiting the author from working full-time on the project and thus the extent of the project is
equivalent to that of 10 ECTS over a 4 month period plus 30 ECTS over a 4 month period.

Sebastian Maklary Jørgensen

___________________________________
Signature

__________________________________
Date 03/08 - 2020

i
Changelog
- V1.0, 3rd of August 2020: Finished report at the date of hand in.
- V2.0, 11th of August 2020: Errors related to multilinear plasticity modeling located and
some results of the simulations added in sections 5.2.6, 5.2.7 and 5.3.
Rules changes regarding loophole added in section 3.1.
- V2.1, 17th of August 2020: Added a comment on the stiffness of multilinear simulation in
section 5.3 along with graph comparisons.
Added comment in section 5.3 on the stiffness of the simulation of the 3PB specimen with
10mm core using the full model with finer mesh.
- V2.2, 20th of August 2020: Added analysis of the PSS test data and comments on an
empirical model to section 5.4
- V2.3, 23rd of September 2020: Name changed on front page due to marriage. Error in
conclusion corrected relating to core material.

ii
Resumé (Danish)
I dette projekt behandles problemstillingen med at modellere laminerede kompositplader med tykke
skumkerner belastet i 3-punkts bøjning. Formålet med modelleringen er at give studerende på
bachelor-niveau de nødvendige værktøjer til at kunne lave foreløbige designvalg i udformningen af
en sikker og regelret monocoque til deltagelse i ingeniørkonkurrencen Formula Student.
Et eksakt bjælkeelement for laminerede bjælker er udledt med inddragelse af
forskydningsdeformation ved antagelse af konstant forskydningstøjning og en korrektionsfaktor
baseret på elasticitetsteori. En 2D bjælkemodel er også benyttet med disse bjælkeelementer som
kompositlagene og QM6 plane elementer som kernemateriale således at kompression af
kernematerialet kan modelleres. Det kommercielle software ANSYS er brugt til at evaluere
bjælkeantagelsen ved sammenligning med shell-elementer og til at lave en 3-dimensionel model
hvor nonlineariteter inkluderes med en bi-lineær plasticitetsmodel. For at validere disse modeller er
fysiske tests af 3-punkts bøjning udført med 5 forskellige lamineringer, så indflydelsen af antallet af
kompositlag, kernetykkelse og kernestivhed kan redegøres for. Materialetests er udført for
kompositmaterialet og for kernematerialet og tests er udført for at udregne stivheden af
testopstillingen så der korrigeres for denne i sammenligningen mellem simulerede og
eksperimentelle data.
De eksperimentelle data af 3-punkts bøjningsemnerne har en spredning på mindre end 2,3% for
stivheden og under 10% for energi absorption og styrker. Sammenlignet med middelværdierne fra
de eksperimentelle resultater er fejlen for stivhed 18,3 til 35,9% for den 1-dimensionelle
bjælkemodel, 2D bjælkemodellen har en fejl på 11,5 til 27,4% og 3D modellen har en fejl på 0,2 til
12,9%. Bjælkemodellerne er ikke generelt retvisende for udregning af styrker og energi absorption
da nonlineriteter ikke medregnes. For 3D modellen er fejlen for energi absorption 6,7 til 55,9%
mens fejlen for styrken er 3 til 35,1% og uoverensstemmelser skyldes antageligvis primært at den
bi-lineære plasticitetsmodel er utilstrækkeligt præcis til at være generelt anvendelig.
Disse modeller er brugt til at finde det optimale laminat til side beskyttelses-strukturen med en vægt
på 6,3217 kg per kvadratmeter, hvilket sandsynligvis resulterer i en monocoque med en vægt under
20 kg.
Penetrationstests som er påkrævede af Formula Student reglerne er også udført, men det er ikke
fundet muligt at modellere styrken af disse emner.

iii
Abstract
In this project, the problem of modelling 3-point bending of laminated composite plates with thick
foam cores are treated with the purpose of enabling bachelor-level students to make preliminary
design choices in regards to the construction of a safe and rules-compliant composite monocoque
for the Formula Student engineering competition. An exact beam element for laminated beams are
developed including axial degree of freedom and bending-elongation coupling due to asymmetry
and shear deformation are accounted for by fist-order shear deformation theory and a correction
factor. A beam model using the mentioned beam elements as skin and QM6 plane elements for the
core is also developed for the inclusion of compressive deformation of the core and the commercial
software ANSYS is used to evaluate the beam assumption by comparison of beam elements to shell
elements and ANSYS is used to create a 3-dimensional nonlinear model for simulation of the full
load-displacement curve with a bilinear plasticity model. To validate these models, physical tests
are conducted with 5 different layups to evaluate the influence of skin thickness, core thickness and
core stiffness. Material tests of the components of the laminate is also conducted to validate the
model against experimental values without relying fully on data sheet data and tests of the testing
machine are conducted to correct for the flexibility of the testing setup.
The experimental data of the 3-point bending specimens are found to have a spread of less than
2,3% for the stiffness and less than 10% for the strengths and energy absorptions. Compared to the
mean of the experimental results, the error of the stiffness is 18,3 to 35,9% for the 1D beam model,
the 2D beam model has errors of 11,5 to 27,4% and the 3D model has an error of 0,2 to 12,9%. The
error for the strengths and energy absorption is great for the 1D and 2D models due to not
modelling nonlinear behavior. For the 3D model, the error of energy absorption is 6,7 to 55,9% and
the error of strength is 3 to 35,1%, with the inconsistencies presumably due to the bilinear plasticity
model being insufficiently accurate.
The models are used to perform a brute-force optimization with regards to mass of the side impact
structure and the lightest rules-compliant laminate has a mass of 6,3217 kg per square meter, which
would likely result in a monocoque mass of less than 20 kg.
Perimeter shear stress tests have also been conducted but attempts at modeling the failures have
failed.

iv
Contents
Preface .................................................................................................................................................. 1
Changelog ............................................................................................................................................ 2
Resumé (Danish) .................................................................................................................................. 3
Abstract ................................................................................................................................................ 4
Contents ............................................................................................................................................... 5
Nomenclature ....................................................................................................................................... 7
List of figures ..................................................................................................................................... 10
List of tables ....................................................................................................................................... 14
1 Introduction .................................................................................................................................. 1
2 Problem statement ........................................................................................................................ 2
3 Theory .......................................................................................................................................... 3
3.1 Rules & regulations ............................................................................................................... 3
3.2 Composite materials .............................................................................................................. 7
3.3 Laminate theory ..................................................................................................................... 9
3.4 Laminated beams ................................................................................................................. 14
3.5 The beam assumption for bending of a plate ...................................................................... 15
3.6 Bernoulli-Euler beams (Classical beam theory) .................................................................. 17
3.7 Timoshenko beams (First-order shear deformation theory) ................................................ 18
3.7.1 Shear correction factor ................................................................................................. 24
3.7.2 Stress calculations ........................................................................................................ 26
3.8 Other beam theories............................................................................................................. 31
3.8.1 Sandwich theory ........................................................................................................... 31
3.8.2 R. P. Shimpi single-variable shear deformation theory ............................................... 32
3.9 2D plane stress finite elements ............................................................................................ 35
3.9.1 Validity of element formulations ................................................................................. 41
3.10 Other 2D and 3D elements for laminate analysis ............................................................ 44
3.11 Nonlinear analysis using ANSYS .................................................................................... 46
3.11.1 Geometric nonlinearities .............................................................................................. 46
3.11.2 Material nonlinearities ................................................................................................. 50
3.11.3 Boundary condition nonlinearities ............................................................................... 54
3.12 Matlab FEM procedure – Boundary conditions .............................................................. 55

v
3.13 Matlab FEM procedure – Solver ..................................................................................... 58
3.14 Statistical analysis ............................................................................................................ 59
4 Method ....................................................................................................................................... 61
4.1 Material choice .................................................................................................................... 61
4.2 Specimen manufacturing ..................................................................................................... 62
4.3 3 point bending test ............................................................................................................. 66
4.3.1 Test setup stiffness correction and baseline tube test .................................................. 68
4.3.2 Simulation method – Matlab ........................................................................................ 71
4.3.3 Simulation method – ANSYS ...................................................................................... 74
4.4 Perimeter shear stress test.................................................................................................... 76
4.4.1 Simulation method – ANSYS ...................................................................................... 77
4.5 Tensile tests ......................................................................................................................... 78
4.6 Compressive tests ................................................................................................................ 78
4.7 Fiber volume fraction analysis ............................................................................................ 79
5 Results ........................................................................................................................................ 82
5.1 Defects ................................................................................................................................. 82
5.1.1 Contaminants ............................................................................................................... 82
5.1.2 Voids ............................................................................................................................ 83
5.1.3 Large-scale surface unevenness ................................................................................... 85
5.2 Material properties .............................................................................................................. 86
5.2.1 Fiber and void volume fraction of SE48LV-HEC ....................................................... 87
5.2.2 Composite tensile test – longitudinal direction ............................................................ 90
5.2.3 Composite tensile test – transverse direction ............................................................... 93
5.2.4 Composite compressive test – longitudinal direction .................................................. 93
5.2.5 Composite compressive test – transverse direction ..................................................... 94
5.2.6 Gurit Corecell M80 tensile test .................................................................................... 96
5.2.7 Gurit Corecell M130 tensile test ................................................................................ 102
5.3 3-point bending.................................................................................................................. 104
5.4 Perimeter shear strength .................................................................................................... 118
5.5 Optimization and recommended layup.............................................................................. 122
6 Conclusions .............................................................................................................................. 126
6.1 Discussion ......................................................................................................................... 127
7 Biblography .............................................................................................................................. 129

vi
Nomenclature
Unless otherwise noted, subscripts denote instances of the variable.
{ } used to denote a column vector
⌊ ⌋ Used to denote a row vector
[ ] used to denote a matrix
[𝑄] constitutive matrix with the components 𝑄𝑖𝑗 , MPa

𝐸 Young’s modulus, GPa


𝐺 Shear modulus, GPa
𝜈 Poissons ratio, unitless
𝐾 Stiffness, N/mm
𝑙 Length, span between supports, mm
𝑡 thickness, mm
ℎ total laminate thickness, mm
𝑏 width, mm
𝐼 Second moment of area
𝐴 cross-sectional area
𝑑𝑖 , 𝑑𝑜 Inner and outer diameter of chassis tube, mm
𝑑 Half of the chassis width, mm
𝑛 counting parameter, unitless
𝓋 volume ratio, unitless
𝓌 weight ratio, unitless
𝒱 volume
𝓂 mass
{𝐹} Force vector with the components 𝐹𝑖 , N or Nm
{𝜎} stress vector with the components 𝜎𝑖 , MPa
𝑓 Tsai-Wu failure criterion factors, MPa-1 or MPa-2
S Strength, MPa
{𝜀} Strain vector with the components 𝜀𝑖 , unitless
[𝑄] constitutive matrix in the lamina coordinate system with components 𝑄𝑖𝑗 , MPa

vii
[𝑄̅ ] constitutive matrix in the global coordinate system with components 𝑄̅𝑖𝑗 , MPa

[𝑇] Transformational rotation matrix for rotating lamina coordinate system 𝜃 degrees into the global
coordinate system, unitless
𝜃 rotation angle from lamina coordinate system into global coordinate system, degrees or radians
𝑢 displacement into the x-direction, mm
𝑣 displacement into the y-direction, mm
𝑤 displacement into the z-direction, mm
𝑢0 , 𝑣0 , 𝑤0 displacements of laminate midplane, mm
𝜙 rotation of cross-section under small angles assumption, equal to the derivative of the deflection
due to bending. Unitless
(0)
{𝜀 (0) } Vector of midplane strain with components 𝜀𝑖
(1)
{𝜀 (1) } Vector of bending curvature with components 𝜀𝑖
{𝑁} Vector of in-plane forces per unit width at the laminate midplane, N/mm
{𝑀} Vector of bending and twisting moments per unit width, Nmm/mm
[𝐴] Lamina midplane elongation stiffness matrix, MPa ⋅ mm
[𝐵] Lamina bending-elongation coupling stiffness matrix, MPa ⋅ mm2
[𝐷] Lamina bending stiffness matrix, MPa ⋅ mm3
[𝐴̅] Lamina midplane elongation stiffness matrix for inverse constitutive relations, (MPa ⋅ mm)-1
[𝐵̅ ] Lamina bending-elongation coupling stiffness matrix for inverse constitutive relations, (MPa ⋅
mm2)-1
̅ ] Lamina bending stiffness matrix for inverse constitutive relations, (MPa ⋅ mm3)-1
[𝐷
𝒬 Through-thickness shear force per unit width, N/mm
[𝒟] Bending stiffness matrix for homogeneous, isotropic plate, MPa ⋅ mm3
𝐷 Flexural stiffness of plate
𝑀 bending moment, Nm
𝑉 Through-thickness shear force, N
Φ Shear stiffness component in Timoshenko beam element, unitless
𝜅 Shear stiffness correction factor
𝐶𝑖 integration constant

viii
𝑙𝑒 Element length, mm
[𝑘] Element stiffness matrix with coefficients 𝑘𝑖𝑗

{𝑑} Vector of the nodal displacements


𝑈 Potential strain energy
𝒩 shape function
𝔅 strain-displacement function
[𝐽] Jacobian matrix.
𝐽 the Jacobian, the determinant of [𝐽]
[Γ] the inverse of [𝐽]
𝜉, 𝜂 Auxiliary coordinates for developing isoparametric elements, unitless.
𝒮𝑖𝑗 components of the second Piola-Kirchoff stress tensor

𝑒 Green-lagrange strains

The Voight-Kelvin notation will be used. 𝜎1 = 𝜎11 , 𝜎2 = 𝜎22 , 𝜎3 = 𝜎33 , 𝜎4 = 𝜎23 , 𝜎5 = 𝜎13 , 𝜎6 =
𝜎12
𝜀1 = 𝜀11 , 𝜀2 = 𝜀22 , 𝜀3 = 𝜀33 , 𝜀4 = 2𝜀23 , 𝜀5 = 2𝜀13 , 𝜀6 = 2𝜀12
At times, the 1-, 2-, and 3-directions are referred to as the x-, y- and z- directions.

ix
List of figures
Figure 1: Different parts of a spaceframe chassis, Viing X taken as example. ................................... 3
Figure 2: Monocoque design based off of the Viking X spaceframe chassis ...................................... 5
Figure 3: Beam with offset coordinate system ................................................................................... 15
Figure 4: Anticlastic curvature seen underneath bent foam plate ...................................................... 16
Figure 5: Anticlastic curvature seen at edges of bent foam plate ...................................................... 17
Figure 6: Timoshenko beam cross-section rotation and centerline slope .......................................... 19
Figure 7: Boundary conditions for timoshenko beam with arbitrary axial displacement at one end. 21
Figure 8: Beam element with an arbitrary transverse displacement at one end ................................. 22
Figure 9: Beam element with an arbitrary rotation of the cross-sectionat one end ........................... 22
Figure 10: Sketch of Timoshenko beam in 3 point bending .............................................................. 26
Figure 11: Through-thickness shear stress of an asymmetric beam .................................................. 31
Figure 12: Comparison of shear stiffness of sandwich and timoshenko theories .............................. 32
Figure 13: Comparison of timoshenko and shimpi beam elements ................................................... 35
Figure 14: Patch test of plane elements ............................................................................................. 41
Figure 15: Plane element beam in 3 point bending ............................................................................ 42
Figure 16: Comparison of plane elements to analytical beam solutions. Stiffness in Newton per
millimeter of vertical deflection of the loaded central node .............................................................. 43
Figure 17: Zoomed-in view of Comparison of plane elements to analytical beam solutions ............ 43
Figure 18: Convergence of refinement of plane element beam ......................................................... 43
Figure 19: Stiffness of plate elements compared to theoretical values. [10] ..................................... 45
Figure 20: Force convergence plot in ANSYS .................................................................................. 49
Figure 21: Schematic view of a bilinear model ................................................................................. 51
Figure 22: Possible bilinear models to fit a certain dataset................................................................ 52
Figure 23: Multilinear hardening model ............................................................................................ 52
Figure 24: Idealized bilinear graph of isotropic and kinematic hardening. ....................................... 53
Figure 25: Calculated stiffness versus penalty stiffness .................................................................... 56
Figure 26: Calcualted stiffness versus penalty stiffness, zoomed in.................................................. 56
Figure 27: Calculated runtime versus element count for different boundary condition methods ...... 57
Figure 28: Runtimes normalized against penalty method runtime versus element count for different
boundary condition methods .............................................................................................................. 57
Figure 29: Runtime versus element count for solving method .......................................................... 58
Figure 30: Ratio of runtime between inversion and mldivide(K,F) versus element count ................ 58
Figure 31: Runtimes versus element count for full matrix and sparse matrix ................................... 59
Figure 32: Comparison of runtime versus element count of full matrix and sparse matrix .............. 59
Figure 33: Comparison of foam core materials from Gurit. Specific shear stiffness has the units
GPa/(kg/m3) while density has the units kg/m3. ................................................................................ 62
Figure 34: Glass plate with tacky-tape and one layer of prepreg ....................................................... 64
Figure 35: Schematic view of application of tacy-tape at a corner .................................................... 64
Figure 36: Laminate within a vacuum bag on a glass plate ............................................................... 65
Figure 37: Drawing of the 3-point bending test fixture ..................................................................... 67
Figure 38: 3-point bending test setup seen from an angle ................................................................. 67
Figure 39: 3-point bending test setup seen from the end ................................................................... 67
Figure 40: Baseline tubes in 3-point bending .................................................................................... 69

x
Figure 41: Local deformation of baseline tubes................................................................................. 69
Figure 42: Baseline tubes 3PB load-displacement graph................................................................... 70
Figure 43: Baseline tubes 3PB stiffness graph................................................................................... 70
Figure 44: Load-displacement graph of 3PB of 10 mm steel plate ................................................... 70
Figure 45: Stiffness graph of 3PB of 10 mm steel plate .................................................................... 70
Figure 46: Singularity in laminate stiffness for SES.......................................................................... 71
Figure 47: Shear stress in core of beam with a 2D plane element core. The two green elements in
the yellow part are a plotting error caused by all of the nodes having the same calculated shear
stress. .................................................................................................................................................. 73
Figure 48: Von Mises equivalent stresses in the core of a beam with 2D plane elements for core ... 74
Figure 49: Boundary conditions of a shell element model (full plate shown) ................................... 75
Figure 50: Full model boundary conditions. Quarter plate shown. Symmetry boundary conditions
are on the faces and edges to simulate a full plate. ............................................................................ 75
Figure 51: Perimeter shear stress test ................................................................................................. 76
Figure 52: Perimeter shear stress test peak detection ........................................................................ 77
Figure 53: ANSYS model of PSS test with solid-shell top skins ...................................................... 78
Figure 54: ASTM D695 compressive test fixture .............................................................................. 79
Figure 55: Cut surface showing the cross-section of a skin consisting of 4 layers of CFRP in
alternating 0 and 90-degree directions ............................................................................................... 82
Figure 56: Contaminants catching the light when seen a bit from the side ....................................... 83
Figure 57: Contaminants in circles. Imprints of wrinkles of the release film seen as horizontal lies
that catches the light ........................................................................................................................... 83
Figure 58: Voids seen when looking parallel to the fibers. The rulers are 2mm and 0,5mm long. ... 84
Figure 59: Voids photographed with microscope built-in camera, looking onto the ends of the
fibers. Voids marked from the left with three 20 micrometers voids and two 10 micrometer voids. 84
Figure 60: Voids looking onto the ends of the fibers, photographed with an external camera through
the eye-piece. Note that the uneven top surface is filled with fibers and are not just pooled resin. . 84
Figure 61: Large void seen parallel to the fibers direction. ............................................................... 85
Figure 62: Large void seen when looking onto the 90-degree direction. .......................................... 85
Figure 63: Large-scale surface unevenness visible on a 3PB specimen ............................................ 86
Figure 64: Void volume fraction plotted against fiber volume fraction ............................................ 89
Figure 65: Nail polish weight plotted against fiber volume fraction ................................................. 89
Figure 66: Specimen density plotted against fiber volume fraction .................................................. 89
Figure 67: Stress-strain diagram of SE84LV-HEC, longitudinal direction ....................................... 91
Figure 68: Stiffness plot of SE84LV-HEC, longitudinal direction .................................................... 92
Figure 69: End-failure of SE84LV-HEC compressive specimen ...................................................... 94
Figure 70: Stress-strain diagram of SE84LV-HEC in compression, transverse direction ................. 95
Figure 71: Stiffness diagram of SE84LV-HEC in compression, transverse direction....................... 95
Figure 72: Fracture of compression specimen under a microscope ................................................... 96
Figure 73: Engineering stress-strain diagram for Corecell M80 tensile test...................................... 97
Figure 74: Stiffness diagram for Corecell M80 tensile test, engineering stress/strain ...................... 97
Figure 75: Zoomed in view of the stiffness diagram of M80 from engineering stress/strain ............ 97
Figure 76: Zoomed in view of the stiffness diagram of M80 from true stress/strain ........................ 97
Figure 77: True stress versus logarithmic strain, M80 tensile test..................................................... 99
Figure 78: True stress versus logarithmic plastic strain, M80 tensile test ......................................... 99

xi
Figure 79: Fit of Ludwik's model onto specimen 2 of M80 tensile test ............................................ 99
Figure 80: Comparison of exprimental and simulated data of 3PB of M80 foam using bilinear
isotropic hardening with experimentally derived material data, Load-Displacement graph ........... 100
Figure 81: Comparison of exprimental and simulated data of 3PB of M80 foam using bilinear
isotropic hardening with experimentally derived material data, Stiffness graph ............................. 100
Figure 82: Load-displacement graph of 3PB of M80 foam. Comparison of experimental data and
simulated with multilinear plasticity model ..................................................................................... 101
Figure 83: True stress-strain diagram of M130 tensile tests ............................................................ 102
Figure 84: Stiffness diagram of M130 tensile test, from true stress and logarithmic strain ............ 102
Figure 85: Load-Displacement graph of 3PB of M130 foam. Comparison of experimental results to
simulated with multilinear plasticity. ............................................................................................... 104
Figure 86: Compression of core seen at the middle of a 3PB test specimen ................................... 104
Figure 87: Load-displacement graph of specimen 1 ........................................................................ 105
Figure 88: Stiffness graph of specimen 1......................................................................................... 105
Figure 89: 3PB load-displacement curve, 4 layer skins, 25mm M80 core ...................................... 106
Figure 90: 3PB stiffness graph, 4 layer skins, 25mm M80 core ...................................................... 106
Figure 91:3PB load-displacement curve, 4 layer skins, 25mm M130 core ..................................... 107
Figure 92:3PB stiffness graph, 4 layer skins, 25mm M130 core ..................................................... 107
Figure 93:3PB load-displacement curve, 4 layer skins, 10 mm M80 core ...................................... 107
Figure 94:3PB stiffness graph, 4 layer skins, 10 mm M80 core ...................................................... 107
Figure 95: 3PB load-displacement graph, 8 layer skins, 25mm M80 core ...................................... 108
Figure 96: 3PB stiffness graph, 8 layer skins, 25mm M80 core ...................................................... 108
Figure 97: 3PB Load-displacement graph, 14 layer skins, 25mm M80 core................................... 108
Figure 98: 3PB stiffness graph, 14 layer skins, 25mm M80 core .................................................... 108
Figure 99: Load-Displacement curve comparison between simualted and tested data, 3PB, specimen
1 ........................................................................................................................................................ 109
Figure 100: Stiffness graph comparison between simulated and tested data, 3PB, specimen 1 ...... 109
Figure 101: Legend for comparison graphs. .................................................................................... 109
Figure 102: Laminate stiffness versus skin thickness ...................................................................... 110
Figure 103: Laminate stiffness versus core thickness ...................................................................... 110
Figure 104: Laminate stiffness versus core density. The densities here are considered the
identification of the foams in the Gurit Corecell M-series .............................................................. 110
Figure 105: Energy absorption versus skin thickness ...................................................................... 110
Figure 106: Energy absorption versus core thickness ...................................................................... 110
Figure 107: Energy absorption versus core density ......................................................................... 110
Figure 108: Yield load versus skin thickness................................................................................... 111
Figure 109: Yield load versus core thickness .................................................................................. 111
Figure 110: Yield load versus core density ...................................................................................... 111
Figure 111: Maximum load versus skin thickness ........................................................................... 111
Figure 112: Maximum load versus core thickness ........................................................................... 111
Figure 113: Maximum load versus core density .............................................................................. 111
Figure 114: Typical failure at first peak of PSS test ........................................................................ 118
Figure 115: Cross-section of PSS test of only M80 foam................................................................ 119
Figure 116: Load-displacement graph of PSS test, 4 layer skin, 25mm M80 core ......................... 119
Figure 117: Load-displacement graph of PSS test, 4 layer skin, 25mm M130 core ....................... 119

xii
Figure 118: Load-displacement graph of PSS test, 4 layer skin, 10 mm M80 core ........................ 120
Figure 119: Load-displacement graph of PSS test, 8 layer skin, 25mm M80 core ......................... 120
Figure 120: Load-displacement graph of PSS test, 14 layer skin, 25mm M80 core ....................... 120
Figure 121: Stiffness of laminates by beam model for optimization purposes ................................ 123
Figure 122: Specific stiffness of laminates by beam model for optimization purposes .................. 123
Figure 123: Stiffness of laminates by 2D beam model for optimization purposes .......................... 124
Figure 124: Specific stiffness of laminates by 2D beam model for optimization purposes ............ 124
Figure 125: Simulation of optimum layup in ANSYS. Elastic shear strain through the core is plotted
.......................................................................................................................................................... 125
Figure 126: Load-displacement graph of optimum layup, simulated by ANSYS full model ......... 125
Figure 127: Stiffness graph of optimum layup, simulated by ANSYS full model .......................... 125

xiii
List of tables
Table 1 Minimum dimensional requirements for baseline chassis. ..................................................... 4
Table 2: Stiffening effect for beams of mediocre width. [7] ............................................................. 15
Table 3: Shape functions for timoshenko beam element ................................................................... 27
Table 4: Strain-displacement functions for laminated Timoshenko beam element ........................... 28
Table 5: Shape functions for quadrilateral elements. [10] ................................................................. 37
Table 6: Derivatives of shape functions from Table 5 ....................................................................... 38
Table 7: Gauss Quadrature weight factors. [5] .................................................................................. 39
Table 8: Material properties of SE84LV-HEC from data sheet......................................................... 86
Table 9: Material properties for Gurit Corecell M80 and M130 from data sheet .............................. 87
Table 10: Material properties of Gurit SE84LV-HEC derived from tests ......................................... 87
Table 11: Material properties of Gurit Corecell M80 and M130 derived from tests ......................... 87
Table 12: Fiber and void volume fraction of SE84LV-HEC ............................................................. 88
Table 13: Hypothesis testing of fiber and void volume fractions ...................................................... 90
Table 14: Stiffness and strength of SE84LV-HEC, longitudinal direction........................................ 93
Table 15: Stiffness and strength of SE84LV-HEC, transverse direction ........................................... 93
Table 16: Stiffness and strength of SE84LV-HEC in compression, transverse direction ................. 96
Table 17: Stiffnesses and strengths of Gurit Corecell M80 foam ...................................................... 98
Table 18: Fitting coefficients for Ludwik's model and bilinear tangent modulus for M80 tensile test
.......................................................................................................................................................... 100
Table 19: Stress levels for multilinear model of M80 foam ............................................................ 101
Table 20: Stiffness and strength of M130 tensile test ...................................................................... 103
Table 21: Coefficients of Ludwik's model and the calculated tangent modulus for M130 foam
tensile tests ....................................................................................................................................... 103
Table 22: Stress and strain values for a multilinear hardening model, M130 tensile tests .............. 103
Table 23: Summary of experimental data, 3-point bending tests .................................................... 112
Table 24: Simulation error, stiffness ................................................................................................ 113
Table 25: Simulation error, energy absorption ................................................................................ 115
Table 26: Simulation error, Yield load ............................................................................................ 115
Table 27: Simulation error, maximum load ..................................................................................... 116
Table 28: Stiffening effect of laminated beam width ...................................................................... 117
Table 29: mean forces for first and second peak of the PSS tests along with % spread .................. 121

xiv
1 Introduction
University students at bachelor and master levels from around the world compete in the engineering
competition FSAE/Formula Student, in which they are required to design and manufacture a single-
seater race car according to a common ruleset. The ruleset regarding chassis structure is built
around the concept of a baseline steel spaceframe but allows for alternative constructions made
from metals or composites if the safety of the alternative structure can be proven to be equivalent to
the baseline.
The team at the University of Southern Denmark, SDU-Vikings, have existed since 2006 and have
built ten cars which have all had a traditional spaceframe but now the team wishes to look into the
possible weight savings of using composites. A monocoque is a chassis structure in which the skin
is the load-bearing structure instead of distinct frame members and such it allows not only for a
reduction of the weight of the load-bearing structure but also for a reduction in the needed
bodywork, which otherwise forms a shell around the frame of a spaceframe chassis. Monocoques
are used both in aerospace, high-level racecars, boats and also roadgoing cars because it poses a
more efficient use of the material and such it has also been used by many Formula Student teams.
Monocoques are often formed as a laminate with a stiff skin on either side of a lightweight and
relatively flexible core in order to raise the bending stiffness and more efficiently use the materials.
The flexibility of the core means that shear deformation cannot be neglected when assessing the
bending stiffness of the plate. Shear deformation of such sandwich plates are not currently taught to
the bachelor-level students whom the team mostly consist of and therefore, this thesis seeks to
present an accurate way of simulating this bending in a way that can be used by the team.

Page 1 af 125
2 Problem statement
It is sought to build a tool which can calculate the stiffness, strength and energy absorption of
laminated composite plates with a light-weight core under 3-point bending. The tool may consist of
simple FEM code in combination with commercial software to gain the capability of both analyzing
large amounts of layups and also get the necessary accuracy to recommend optimum layups to be
manufactured for testing. The tool should be sufficiently simple that it is usable by bachelor-level
engineering students as the team SDU-Vikings primarily consists of these. To verify the accuracy of
the calculations, physical tests of laminates must be conducted.
A method to determine the load required to penetrate the composite is also sought to enable
simulation of the perimeter shear stress test as well.

Objectives:
To develop a finite element model using beam elements, capable of approximately calculating the
stiffness, strength and energy absorption of laminated composite plates with a light-weight core.
To set up a finite element model using commercial software, capable of accurately calculating the
stiffness, strength and energy absorption of laminated composite plates with a light-weight core.
To validate the finite element models by physical tests in which multiple laminate parameters are
assessed. The parameters varied are the amount of carbon fiber layers, the thickness of the core and
the core stiffness.
To use the developed model as a preliminary design tool to search for the lightest rules-compliant
laminates that can be used to construct a composite monocoque.

Page 2 af 125
3 Theory

3.1 Rules & regulations


The Formula Student rules are organized as an international ruleset with competition-specific
supplementary rules and several templates for required documentation. Section T3 of the Formula
Student rules 2020 defines the requirement for the chassis with T3.2 giving the minimum
requirements for a baseline chassis and T3.3 to T3.5 defining how alternative materials can be used
and equivalency should be proven through testing. An excel worksheet called the Structural
Equivalency Spreadsheet, abbreviated SES, provides a template to proving the chassis to be rules
compliant. [1]
The chassis is divided into sections which have different requirements of stiffness, energy
absorption, strength and method of calculation. For all chassis sections, it is allowed to make a
spaceframe chassis which deviates from the dimensions of the baseline chassis tubes if the tubes
that are used exceed the minimum wall-thickness, cross-sectional area and area moment of inertia in
Table 1. The requirements of spaceframes using alternative materials are not further discussed here.
On Figure 1, The chassis of Viking X is seen in side-view and the tubes that make up different
sections of the chassis are marked by numbers, which are put in parentheses in Table 1 to show
which parts of the chassis is named what. Some parts of the chassis may have multiple functions
such as the tubes that protect the tractive system and the main hoop bracing supports, in which case
the highest requirement must be met.

Figure 1: Different parts of a spaceframe chassis, Viing X taken as example.

Page 3 af 125
Item or application Minimum Minimum Minimum area
wall cross- moment of inertia
thickness sectional area
Main hoop (1) 2,0 [𝑚𝑚] 175 [𝑚𝑚2 ] 11320 [𝑚𝑚4 ]
Front hoop (2)
Shoulder harness mounting bar (8)
Side impact structure (6) 1,2 [𝑚𝑚] 119 [𝑚𝑚2 ] 8509 [𝑚𝑚4 ]
Front bulkhead (4)
Roll hoop bracings (9) (3)
Driver’s harness attachment (except shoulder
harness bar)
Accumulator container protection structure (7)
Tractive system protection structure (7)
Front bulkhead support (5) 1,2 [𝑚𝑚] 91 [𝑚𝑚2 ] 6695 [𝑚𝑚4 ]
Main hoop bracing support (7)
Table 1 Minimum dimensional requirements for baseline chassis.

The main roll hoop and its bracing must be made from steel tubing and the front hoop must be made
from metallic tube but the rest of the chassis is free in its design and materials as long as
equivalency is proven using the structural equivalency spreadsheet. If the spaceframe chassis of the
Viking X is made into a monocoque by removing the chassis tubes, making planes between the
chassis nodes and assigning a laminate thickness to those planes, the monocoque might look
something like Figure 2.

Page 4 af 125
Figure 2: Monocoque design based off of the Viking X spaceframe chassis

For all laminated parts of the monocoque, the cross-sectional area of the skins must exceed the
minimum cross-sectional area stated in Table 1 for the tubes that the laminated part replaces
compared to a baseline chassis. Also, for all laminated parts, the bending stiffness must exceed that
of the tubes that it replaces. In this regard, the bending stiffness is calculated as the second moment
of area, 𝐼𝑐 , of the composite skins multiplied by an equivalent skin stiffness 𝐸𝑒𝑞 which is calculated
from 3 point bending of a representative test specimen. The SES calculates the equivalent skin
stiffness as in equation (3-1) where 𝐾𝑙𝑎𝑚 is the corrected stiffness of the test specimen, 𝑙 is the test
span and 𝐼𝑙𝑎𝑚 is the second moment of area of the tested specimen. The second moment of area are
calculated per equation (3-2) in which, 𝑏 is the specimen width and 𝑡 is the thickness of each part of
the laminate.

−3
𝐾𝑙𝑎𝑚 ⋅ 𝑙 3 (3-1)
𝐸𝑒𝑞 = 10 ⋅
48 ⋅ 𝐼𝑙𝑎𝑚

1 (3-2)
𝐼𝑙𝑎𝑚 = 𝑏 ⋅ ((𝑡𝑐𝑜𝑟𝑒 + 𝑡𝑖𝑛𝑛𝑒𝑟𝑠𝑘𝑖𝑛 + 𝑡𝑜𝑢𝑡𝑒𝑟𝑠𝑘𝑖𝑛 )3 − 𝑡𝑐𝑜𝑟𝑒
3 )⋅
12
The second moment of area of the skins of the actual laminate, 𝐼𝑐 , is calculated with the parallel
axis theorem under the assumption that the laminate is a flat panel and with these values, the SES
calculates the laminate as a beam consisting only of the skins and without shearing deformation. For
all sections of the chassis but the side impact structure, the calculated values 𝐸𝑐 𝐼𝑐 and the strength

Page 5 af 125
results from the 3-point bending test are used to calculate theoretical values of tensile strength, max
load at midspan of a 1m beam, max deflection of a 1m beam and the energy absorption of elastic
deformation of a beam until the ultimate tensile strength.

The Side Impact Structure, abbreviated SIS, is the sides and floor of the cockpit between the two
roll hoops and is subject to the strictest requirements as the entire driver, except the legs, is
protected by this section of chassis. The side of the SIS must be equivalent to two chassis tubes
while half the floor must be equivalent to a third tube. This equivalency is only calculated
theoretically for the bending stiffness and the cross-sectional area, which are both calculated under
a flat panel assumption regardless of the curvature of the structure.
The remaining equivalency requirements are taken directly from the tested specimen. The highest
tested force and the absorbed energy of the tested specimen for the side must exceed the values
from the test of the two baseline tubes it replaces regardless of the dimensions of the actual
structure. Additionally, the laminate used for the SIS requires that the highest peak of a perimeter
shear stress test to be at least 7,5 kN.
When proving equivalency of the main hoop bracing support (MHBS), the front hoop bracing
(FHB) and the front bulkhead support structure (FBHS), the actual shape of the car can be taken
into account to show equivalency if the flat-panel calculations show between 50% and 100%
equivalency. The way the actual shape of the car is taken into account for the MHBS and the BFHS
is to compare the second moment of area with respect to the car centerline against a corresponding
value for a baseline chassis. The second moment of area of the laminate is found by making a cross-
section in the CAD drawing and calculating the combined second moment of area of the skins of
one half of the car. The corresponding value for a baseline chassis is such that with 𝑑, 𝑑𝑖 and 𝑑𝑜
being respectively half the chassis width and the inner and outer diameter of the 𝑛𝑡 baseline tubes
then the second moment of area, 𝐼, of the baseline chassis is calculated by the parallel axis theorem
in equation (3-3).
𝜋 4 𝜋 (3-3)
𝐼 = 𝑛𝑡 ⋅ (𝑑𝑜 − 𝑑𝑖4 ) + 𝑛𝑡 ⋅ (𝑑𝑜2 − 𝑑𝑖2 ) ⋅ 𝑑 2
64 4
For accounting for the actual form of the front hoop bracings, the second moment of area of the
baseline tubes and of the composite structure are used without referring them to the car centerline
using the parallel axis theorem.
The front bulkhead support structure must be constructed of laminate which has been tested to at
least 4 kN in the perimeter shear strength test.
If bolted joints are used in the assembly of the chassis or attachment of the harness to it, rule T 3.16
requires a load carrying capability of 30 kN, which is to be ensured by extrapolating the area from
the first peak of the perimeter shear strength tests.
Other significant rules include a requirement that at most 50% of the fiber weight are within +/- 10
degrees and that it is permitted to use one laminate for testing and another for the construction of
the monocoque only if nothing but the thickness of the core differs between the laminates. The
latter rule can be considered a loophole as the calculated 𝐸𝑐 will be much higher with a thin core

Page 6 af 125
than with a thick core due to the assumption of it being a beam with no shear. This loophole will not
be exploited in this paper.
The following in this subsection is added in V2.0 of this report, the 11th of august 2020
The loophole regarding the impact of core thickness on the calculated 𝐸𝑐 has been closed in the
2020 SES version 1.1.2.
A guidance note has been added which specifies that the rule about testing with one core thickness
and using another is only applicable if the tested laminate from which 𝐸𝑐 is derived has a thicker
core than that which is actually being used.

3.2 Composite materials


Fiber-reinforced composite materials consist of fibers embedded in a matrix where the fibers are
intended to be the main load-carrying component while the matrix supports the fibers, transfers load
between them and protects them from the environment. As the individual fibers are approximately
one-dimensional and are oriented primarily in one direction, the material properties are highly
orthotropic as opposed to isotropic materials such as steel or aluminum which don’t have a
preferred direction.
The elastic properties of a fiber reinforced polymer can be calculated from the properties of the
fiber and matrix or it can be determined from tests. For a unidirectional lamina, the 1-direction is
defined as the fiber longitudinal direction, the 2-direction is defined as the transverse in-plane
direction and the 3-direction is defined as the out of plane direction and the elastic properties are
calculated as in equations (3-4) to (3-10), stated in [2].
The elastic modulus in the 1-direction and the Poisson’s ratio in the 6-direction are calculated using
the rule of mixtures, which assumes that the fiber and matrix each act in parallel of each other while
the elastic modulus in the transverse direction and the in-plane shear modulus is calculated from a
model where the fibers and matrix are assumed to act in series of each other. Because of transverse
isotropy, the number of independent elastic properties are reduced as some of them are the same.
Subscript 𝑓 denotes fiber properties and subscript 𝑚 denotes matrix properties.
𝐸1 = 𝐸𝑓 𝑣𝑓 + 𝐸𝑚 𝑣𝑚 (3-4)

𝐸𝑓 𝐸𝑚 (3-5)
𝐸2 = 𝐸3 =
𝐸𝑓 𝓋𝑚 + 𝐸𝑚 𝓋𝑓
𝐺𝑓 𝐺𝑚 (3-6)
𝐺6 = 𝐺5 =
𝐺𝑓 𝓋𝑚 + 𝐺𝑚 𝓋𝑓
𝐸2 (3-7)
𝐺4 =
2(1 + 𝜈23 )
𝜈12 = 𝜈13 = 𝜈𝑓 𝓋𝑓 + 𝜈𝑚 𝓋𝑚 (3-8)

𝐸2 (3-9)
𝜈21 = 𝜈31 = 𝜈12
𝐸1
1 − 𝜈21 (3-10)
𝜈23 = 𝜈32 = 𝜈12
1 − 𝜈12

Page 7 af 125
Like the stiffness is orthotropic, so is the strength as this also depends on the orientation. Because of
the nonhomogeneous nature of the composite, the loads to failure vary greatly with the direction of
loading such that the tensile strength is often much higher than the compressive strength because
the failure modes are different. With tension along the fiber direction, the failure is often caused by
fiber fracture as the fibers have a much lower elongation-to-failure than the matrix but in
compression along this direction, the failure is often caused by buckling of the fibers within the
matrix. In the transverse direction, the strength is limited by the strength of the matrix but is further
reduced from this due to voids and that the fibers act as solid inclusions, both of which are creating
stress concentrations within the matrix.
The in-plane shear failure is caused by a failure of the matrix or the matrix-fiber interface and while
each lamina expresses transverse isotropy, the transverse shear strength is often limited by the
interlaminar shear strength. The interlaminar shear strength is the shear strength of the interface
between layers where there may be a lower fiber volume fraction due to fibers crossing at an angle
or there may be a higher concentration of voids as discussed in the defects chapter of this report.
Failure prediction is done either by failure theories, by means of fracture mechanical analysis or by
a combination. To limit the scope of this project, fracture mechanical analysis is omitted. Unlike
ductile homogeneous materials where Von Mises yield criterion is considered almost universally
applicable, there is no single widely accepted failure criterion. Instead, many competing theories of
varying complexity are assessed. Most common failure theories are maximum strain theory, max
stress theory, Azzi-Tsai-Hill failure criterion, Tsai-Wu failure criterion and Hashin failure criterion.
The maximum stress theory does not account for combination of stresses but state that failure will
occur if the stresses in any of the principal material directions exceed the strength in that direction.
Maximum strain theory is similar but predicts failure if the strain exceeds the failure strain in any of
those directions.
The Azzi-Tsai-Hill theory accounts for combination of stresses by predicting failure of an
orthotropic lamina in plane stress condition if equation (3-11) is fulfilled. 𝑆1 and 𝑆2 are taken as
either the tensile or compressive strength depending on the sign of the corresponding stress. [2]
𝜎12 𝜎1 𝜎2 𝜎22 𝜎62 (3-11)
− 2 + 2+ 2=1
𝑆12 𝑆1 𝑆2 𝑆6
The Tsai-Wu failure criterion states that under plane stress, the orthotropic lamina will fail if
equation (3-12) is true. [2]
2
𝑓1 𝜎1 + 𝑓2 𝜎2 + 𝑓6 𝜎6 + 𝑓11 𝜎11 + 𝑓22 𝜎22 + 𝑓66 𝜎62 + 2𝑓12 𝜎1 𝜎2 = 1 (3-12)

Where the factors except 𝑓12 are given in equation (3-13). The factor 𝑓12 must be determined by a
biaxial test which are rarely done and thus the value is often approximated as the lower limit of
equation (3-14). The Tsai-Wu criterion is criticized for not having the value of 𝑓12 uniquely defined
from biaxial tests as it may be found to be different depending on whether the biaxial test is done
with compressive or tensile stresses and also, the criterion is criticized for having the prediction of a
tensile failure depend on the compressive strength. [3]

Page 8 af 125
1 1 1 1
𝑓1 = − , 𝑓2 = − ,𝑓 = 0
𝑆1𝑡 𝑆1𝑐 𝑆2𝑡 𝑆2𝑐 6 (3-13)

1 1 1
𝑓11 = , 𝑓22 = , 𝑓66 = 2
𝑆1𝑡 𝑆1𝑐 𝑆2𝑡 𝑆2𝑐 𝑆6
1 1 (3-14)
− (𝑓11 𝑓22 )2 ≤ 𝑓12 ≤ 0
2
The Hashin failure criterion is a proposed solution to the criticism of the Tsai-Wu theory and does
not claim that the failure can be predicted by a single smooth curve. The Hashin failure criterion is a
set of four equations, each accounting for a failure mode such that the failure of fibers and matrix in
both tension and compression are considered separately. In plane stress, equations (3-15) to (3-18)
describe the failure criterion.
𝜎1 2 𝜎6 2 (3-15)
( ) + ( ) = 1 𝑓𝑜𝑟 𝜎1 > 0
𝑆1𝑡 𝑆6
𝜎1 2 (3-16)
( ) = 1 𝑓𝑜𝑟 𝜎1 < 0
𝑆1𝑐
𝜎2 2 𝜎6 2 (3-17)
( ) + ( ) = 1 𝑓𝑜𝑟 𝜎2 > 0
𝑆2𝑡 𝑆6

𝜎2 2 𝑆2𝑐 2 𝜎2 𝜎6 2 (3-18)
( ) + (( ) − 1) + ( ) = 1 𝑓𝑜𝑟 𝜎2 < 0
2𝑆4 𝑆4 𝑆2𝑐 𝑆6

These failure criteria also exist for 3-dimensional stress states but due to the availability of material
properties from data sheets and simple tests, these are not included in the present paper.

3.3 Laminate theory


The laminated plate consists of several specially orthotropic lamina which have the stress-strain
relation in equation (3-19) with the components given in equation(3-20), taken from [4].
𝜎1 𝑄11 𝑄12 𝑄13 0 0 0 𝜀1 (3-19)
𝜎2 𝑄12 𝑄22 𝑄23 0 0 0 𝜀 2
𝜎3 𝑄13 𝑄23 𝑄33 0 0 0 𝜀3
𝜎4 = 0 0 0 𝑄44 0 0 𝜀4
𝜎5 0 0 0 0 𝑄55 0 𝜀5
{𝜎6 } [ 0 0 0 0 0 𝑄66 ] 𝜀6 }
{

Page 9 af 125
where 1 − 𝜈23 𝜈32 𝜈21 + 𝜈31 𝜈23 𝜈12 + 𝜈32 𝜈13 𝜈31 + 𝜈21 𝜈32 (3-20)
𝑄11 = , 𝑄12 = = , 𝑄13 =
𝐸2 𝐸3 ∆ 𝐸2 𝐸3 ∆ 𝐸1 𝐸3 ∆ 𝐸2 𝐸3 ∆
𝜈13 + 𝜈12 𝜈23
=
𝐸1 𝐸2 ∆
1 − 𝜈13 𝜈31 𝜈32 + 𝜈12 𝜈31 𝜈23 + 𝜈21 𝜈3 1 − 𝜈12 𝜈21
𝑄22 = , 𝑄23 = = , 𝑄33 =
𝐸1 𝐸3 ∆ 𝐸1 𝐸3 ∆ 𝐸1 𝐸2 ∆ 𝐸2 𝐸2 ∆
𝑄44 = 𝐺4 , 𝑄55 = 𝐺5 , 𝑄66 = 𝐺6
1 − 𝜈12 𝜈21 − 𝜈23 𝜈32 − 𝜈31 𝜈13 − 2𝜈21 𝜈32 𝜈13
∆=
𝐸1 𝐸2 𝐸3
As the lamina may be rotated some angle 𝜃 in the plane of the lamina, the stiffness matrix for the
generally orthotropic is made by the use of the specially orthotropic stiffness matrix [𝑄] and a
transformation matrix [𝑇]
[𝑄̅ ] = [𝑇][𝑄][𝑇]𝑇 (3-21)

𝑄̅11 𝑄̅12 𝑄̅13 0 0 𝑄̅16


𝑄̅12 𝑄̅22 𝑄̅23 0 0 𝑄̅26
𝑄̅13 𝑄̅23 𝑄̅33 0 0 𝑄̅36
=
0 0 0 𝑄̅44 𝑄̅45 0
0 0 0 𝑄̅45 𝑄̅55 0
̅
[𝑄16 ̅
𝑄26 ̅
𝑄36 0 0 ̅
𝑄66 ]
Where
cos 2 (𝜃) sin2 (𝜃) 0 0 0 −2 sin(𝜃) cos(𝜃) (3-22)

sin2 (𝜃) cos2 (𝜃) 0 0 0 2 sin(𝜃) cos(𝜃)


0 0 1 0 0 0
[𝑇] =
0 0 0 cos(𝜃) − sin(𝜃) 0
0 0 0 sin(𝜃) cos(𝜃) 0
2 (𝜃)
[sin(𝜃) cos(𝜃) − sin(𝜃) cos(𝜃) 0 0 0 cos − sin2(𝜃)]

When normal stress in the 3-direction is neglected, which it often is for plates and beams, the
constitutive relation reduces to equation (3-23).
𝜎1 𝑄11 𝑄12 0 0 0 𝜀1 (3-23)
𝜎2 𝑄21 𝑄22 0 0 0 𝜀2
𝜎4 = 0 0 𝑄44 0 0 𝜀4
𝜎5 0 0 0 𝑄55 0 𝜀5
𝜎
{ 6} [ 0 0 0 0 𝑄66 ] {𝜀6 }

Where

Page 10 af 125
𝐸11 𝐸22 (3-24)
𝑄11 = , 𝑄22 =
1 − 𝜈12 𝜈21 1 − 𝜈21 𝜈12
𝜈12 𝐸22 𝜈21 𝐸11
𝑄12 = 𝑄21 = =
1 − 𝜈21 𝜈12 1 − 𝜈12 𝜈21
𝑄66 = 𝐺12 , 𝑄44 = 𝐺13 , 𝑄55 = 𝐺23

The transformation matrix reduces in a similar manner as the rows and columns related to the 3-
direction are removed.
When the transformed lamina constitutive matrix is formed, the formation of a laminate constitutive
matrix requires displacement assumptions. The transverse displacement of the plate is commonly
assumed to not vary through the thickness and such, the transverse displacement of any point within
the laminate is equal to that of the midplane. The in-plane displacements are assumed differently
depending on the used plate theory but generally, they exist as the midplane displacements and one
or more terms which vary through the thickness. For the classical plate theory, also called Kirchoff-
Love plate theory, the variation through the thickness is linearly related to the first derivative of the
transverse deflection such that it is given by equation (3-25) while for the first-order shear
deformation theory (FSDT), the variation through the thickness are linearly related to the rotation of
the cross-sections about the x- and y-axes and the deformation field is given by (3-26). [5]
𝜕𝑤0 (3-25)
𝑢(𝑥, 𝑦, 𝑧) = 𝑢0 (𝑥, 𝑦) − 𝑧
𝜕𝑥
𝜕𝑤0
𝑣(𝑥, 𝑦, 𝑧) = 𝑣0 (𝑥, 𝑦) − 𝑧
𝜕𝑦
𝑤(𝑥, 𝑦, 𝑧) = 𝑤0 (𝑥, 𝑦)

𝑢(𝑥, 𝑦, 𝑧) = 𝑢0 (𝑥, 𝑦) + 𝑧𝜙𝑥 (𝑥, 𝑦) (3-26)

𝑣(𝑥, 𝑦, 𝑧) = 𝑣0 (𝑥, 𝑦) + 𝑧𝜙𝑦 (𝑥, 𝑦)

𝑤(𝑥, 𝑦, 𝑧) = 𝑤0 (𝑥, 𝑦)
Kirchoff-Love plates are analogous to Bernoulli-Euler beams in that shear deformation is assumed
negligible while first-order shear deformation theory is analogous to Timoshenko beams in that
shear deformation is included by a shear strain that is constant through the thickness and shear
correction coefficients are used to correct the shear stiffness. The Kirchoff-love plate theory exist as
a special case of the first-order shear deformation theory as the shear deformation becomes
𝜕𝑤0 𝜕𝑤0
negligible for small thicknesses, leading to 𝜙𝑥 (𝑥, 𝑦) = − and 𝜙𝑦 (𝑥, 𝑦) = − . In this project,
𝜕𝑥 𝜕𝑦
shear deformation is not negligible due to the thick core, so the following equations refer to the first
order shear deformation theory.
The strains are defined in [6] by derivatives of the displacement field and thus, the strains are

Page 11 af 125
𝑑𝑢 𝑑𝑢0 𝑑𝜙𝑥 (3-27)
𝜀1 = = +𝑧
𝑑𝑥 𝑑𝑥 𝑑𝑥
𝑑𝑣 𝑑𝑣0 𝑑𝜙𝑦
𝜀2 = = +𝑧
𝑑𝑦 𝑑𝑦 𝑑𝑦
𝑑𝑤
𝜀3 = =0
𝑑𝑧
𝑑𝑣 𝑑𝑤 𝑑𝑤0
𝜀4 = + = 𝜙𝑦 (𝑥, 𝑦) +
𝑑𝑧 𝑑𝑦 𝑑𝑦
𝑑𝑢 𝑑𝑤 𝑑𝑤0
𝜀5 = + = 𝜙𝑥 (𝑥, 𝑦) +
𝑑𝑧 𝑑𝑥 𝑑𝑥
𝑑𝑢 𝑑𝑣 𝑑𝑢0 𝑑𝑣0 𝑑𝜙𝑥 𝑑𝜙𝑦
𝜀6 = + = + +𝑧 ( + )
𝑑𝑦 𝑑𝑥 𝑑𝑦 𝑑𝑥 𝑑𝑦 𝑑𝑥

The strains 𝜀1 , 𝜀2 and 𝜀6 contain both constant terms and terms which vary linearly through the
thickness so the strains can be expressed as equation (3-28) or in short (3-29).
𝑑𝑢0 𝑑𝜙𝑥 (3-28)

𝜀1
(0)
𝜀1
(1) 𝑑𝑥 𝑑𝑥
𝜀1 𝑑𝑣0 𝑑𝜙 𝑦
{𝜀2 } = {𝜀2(0) } + 𝑧 {𝜀2(1) } = +𝑧
𝜀6 𝑑𝑦 𝑑𝑦
(0) (1)
𝜀6 𝜀6 𝑑𝑢0 𝑑𝑣0 𝑑𝜙𝑥 𝑑𝜙𝑦
+ +
{ 𝑑𝑦 𝑑𝑥 } { 𝑑𝑦 𝑑𝑥 }
{𝜀} = {𝜀 (0) } + 𝑧{𝜀 (1) } (3-29)

By integrating the stresses over the cross-section, the in-plane forces per unit width are calculated in
(3-30) and the bending moment per unit width are calculated in (3-31).
ℎ ℎ (3-30)
𝑧= 𝑧=
2 2
{𝑁} = ∫ 𝜎 𝑑𝑧 =∫ [𝑄̅ ]({𝜀 (0)
} + 𝑧{𝜀 (1)
})𝑑𝑧
ℎ 𝑥𝑥 ℎ
𝑧=− 𝑧=−
2 2
ℎ ℎ (3-31)
𝑧= 𝑧=
2 2
{𝑀} = ∫ 𝑧𝜎𝑥𝑥 𝑑𝑧 = ∫ [𝑄̅ ]𝑧({𝜀 (0) } + 𝑧{𝜀 (1) })𝑑𝑧
ℎ ℎ
𝑧=− 𝑧=−
2 2

To write the laminate constitutive equations, the A-, B- and D-matrices are defined by equation
(3-32), (3-33) and (3-34) and with these, the coupling between strains and applied loads are given
by (3-40) where only the 1-, 2- and 6-directions are included. The coupling between shear strains
and shear deformations are also from the A-matrix because the shear strain doesn’t vary through the
thickness but it is given separately in (3-43) because they don’t interact with the other strains and
(3-40) are how the constitutive relations are given for plates in [2] where shear deformations are
neglected.

Page 12 af 125
ℎ 𝑘 (3-32)
𝑧=
2
[𝐴] = ∫ [𝑄̅ ]𝑑𝑧 = ∑[𝑄̅ ]𝑘 ⋅ (𝑧𝑘 − 𝑧𝑘−1 )

𝑧=−
2
ℎ 𝑘 (3-33)
𝑧=
2 1
[𝐵] = ∫ [𝑄̅ ]𝑧𝑑𝑧 = ∑[𝑄̅ ]𝑘 ⋅ (𝑧𝑘2 − 𝑧𝑘−1
2
)
𝑧=−
ℎ 2
2
ℎ 𝑘 (3-34)
𝑧=
2 1
[𝐷] = ∫ [𝑄̅ ]𝑧 2 𝑑𝑧 = ∑[𝑄̅ ]𝑘 ⋅ (𝑧𝑘3 − 𝑧𝑘−1
3
)
𝑧=−
ℎ 3
2

{𝑁} [𝐴] [𝐵] {𝜀 (0) } (3-35)


{ }=[ ]{ }
{𝑀} [𝐵] [𝐷] {𝜀 (1) }

𝒬𝑦 𝐴 𝐴45 𝜀4 (3-36)
{ } = [ 44 ]{ }
𝒬𝑥 𝐴45 𝐴55 𝜀5

The strains can also be calculated from the applied loads by inversion of the constitutive matrix
using equations (3-44) and (3-47).

{𝜀 (0) } = [𝐴̅]{𝑁} + [𝐵̅ ]{𝑀} (3-37)

{𝜀 (1) } = [𝐶̅ ]{𝑁} + [𝐷


̅ ]{𝑀} (3-38)

here
[𝐴̅] = [𝐴−1 ] + [𝐴−1 ][𝐵][(𝐷⋆ )−1][𝐵][𝐴−1 ] (3-39)

[𝐵̅ ] = −[𝐴−1 ][𝐵][(𝐷⋆ )−1 ]


[𝐶̅ ] = −[(𝐷⋆ )−1 ][𝐵][𝐴−1 ] = [𝐵1 ]𝑇
[𝐷⋆ ] = [𝐷] − [𝐵][𝐴−1 ][𝐵]
̅ ] = [(𝐷⋆ )−1 ]
[𝐷

Within this inversion, the two-dimensional stiffening effect is removed which is most easily shown
by assessing the stiffness of a homogeneous plate.
For a homogeneous, isotropic plate, the bending stiffness matrix is given by equation (3-40).
𝐷 𝐷𝜈 0 (3-40)
𝐷𝜈 𝐷 0
[𝒟] = [ (1 − 𝜈)𝐷]
0 0
2

Page 13 af 125
𝐸𝑡 3 𝐸
With flexural stiffness 𝐷 = 12(1−𝜈2). Here, (1−𝜈2) is the stiffness from the [𝑄]-matrix where the
𝐼
division by (1 − 𝜈 2 ) is a stiffening effect of the 2-dimensional stress state and the rest is simply 𝑏
because plates are normalized to forces/moments per unit width.
When the inverse of this bending stiffness matrix is taken, the first element of the inverted matrix
−1 1 1 12 𝐸𝐼 −1
becomes[𝒟]11 = − 𝐷(𝜈2−1) = 𝐷(1−𝜈2) = 𝐸𝑡 3 = ( 𝑏 ) . As is seen the stiffening effect is removed
−1 )−1
from the first term by the inversion and 𝑏 ⋅ ([𝒟]11 can be used in place of 𝐸𝐼 in beam
calculations.

3.4 Laminated beams


Plates, which lamination theory are most often applied to, are taken to have transverse deflection
into the z-direction while beams are most often taken to have their transverse deflection into the y-
direction. While the orientation of a beam or plate is completely arbitrary, the distinction is
important because a positive moment results in a negative curvature of the transverse displacement
in the z-direction while it causes a positive curvature of the transverse displacement in the y-
𝑑2 𝑣
direction. That is, the bending moment of a homogeneous Bernoulli-Euler beam equals 𝐸𝐼 𝑑𝑥 2
𝑑2 𝑣
when the deflection is in the y-direction while it equals −𝐸𝐼 𝑑𝑥 2 when the deflection is in the z-
direction. In the following, beam elements are developed as having a transverse deflection into the
y-direction so that the formulation of the stiffness matrix is similar to what students are used to for
(1)
beams in the xy-plane. Then the bending strain is taken as the rotation of the cross-section, 𝜀1 =
𝑑𝑣
𝜙𝑥 (𝑥, 𝑦), which approaches the slope of the beam centerline, 𝑑𝑥, when the beam becomes slender.
The method of deriving the A, B and D matrices is still valid, taken that the material properties are
calculated as laminated through the y-direction instead of the z-direction, but with the sign of the
terms with [𝐵]-, [𝐵̅ ]- and [𝐶̅ ]-matrices switched in equation (3-35), (3-37) and (3-38).
To show why the elongation-bending coupling terms must be given the opposite sign, observe on
Figure 3 a beam where the coordinate system does not coincide with the neutral axis when bending.
The “midplane” strain {𝜀 (0) } is just the strain measured at 𝑦 = 0 (𝑧 = 0 for plates) so if this beam is
bent to a positive curvature, the point at 𝑦 = 0 would move outwards unless acted upon by an
external force. If the only external forces are the moment M and the in-plane force N, and the
moment is causing the positive bending curvature, then the force N must be negative to hold the
midplane strain at zero, which requires the coupling terms of the mentioned equations to be
negative. Similar observations can be made if the midplane strain is positive while the curvature is
zero or if only one of M or N are non-zero to show that all of the coupling terms must have the
opposite sign of when the deflection are taken in the z-direction.

Page 14 af 125
Figure 3: Beam with offset coordinate system

3.5 The beam assumption for bending of a plate


The 3-point bending of a plate seems a lot like the bending of a beam with the only differences
being a significant difference in the dimensions of the specimen and the fact that beams have a
bending stiffness that is proportional to the young’s modulus 𝐸, while plates have a flexural
𝐸
stiffness that is proportional to (1−𝜈2). In engineering literature, it is suggested that the bending of a
beam with relatively great widths can be calculated using beam equations but where 𝐸 is replaced
𝐸
by (1−𝜈2). [7]
It is however also stated that the stiffening effect depends on the width, thickness, bending
curvature and poissons ratio through the values given in Table 2 such that intermediate values
1
between 1 and 1−𝜈2 can be calculated for homogeneous beams of mediocre width.

Value of 𝑏 2 𝑑 2 𝑣𝑏
𝜈 𝑡 𝑑𝑥 2
𝟎, 𝟐𝟓 𝟏 𝟒 𝟏𝟔 𝟓𝟎 𝟐𝟎𝟎 𝟖𝟎𝟎
𝟎, 𝟏 1 1,0003 1,0033 1,0073 1,0085 1,0093 1,0097
𝟎, 𝟐 1,0001 1,0013 1,0135 1,0300 1,0349 1,0383 1,0400
𝟎, 𝟑 1,0002 1,0029 1,0311 1,0710 1,0826 1,0907 1,0948
𝟎, 𝟑𝟑𝟑𝟑 1,0002 1,0036 1,0387 1,0895 1,1042 1,1146 1,1198
𝟎, 𝟒 1,0003 1,0052 1,0569 1,1357 1,1584 1,1744 1,1825
𝟎, 𝟓 1,0005 1,0081 1,0923 1,2351 1,2755 1,3045 1,3189
Table 2: Stiffening effect for beams of mediocre width. [7]

These figures come from an analytical analysis of the anticlastic curvature of beams of varying
widths [8]. The anticlastic curvature of a beam is a curvature in the lateral direction which forms the

Page 15 af 125
beam into the shape of a saddle instead of a cylinder and is caused by the Poisson’s ratio which
makes the beam wider on the compression side of the beam and narrower at the tension side of the
beam. For beams that are wide, this anticlastic curvature is suppressed, thus leading to a stiffening
1
effect which for an infinitely wide beam equals 1−𝜈2 as the width goes to infinity. During the
bending of a 10 mm thick, 275mm wide plate, anticlastic curvature can be seen on Figure 4 and
Figure 5 where the curvature is seen respectively at the edges underneath the specimen and is
visible as an air gap between the load applicator and the specimen along the upper side of the edges
of the specimen.

Figure 4: Anticlastic curvature seen underneath bent foam plate

To approximate the values of the stiffening effect, the foam plate shown above and a steel plate in
𝑏 2 𝑑2 𝑣𝑏
the same dimensions are used to calculate values of to compare with Table 2.
𝑡 𝑑𝑥 2

At a load of 16000 Newton, tested at a span of 400 mm, the steel plate is approximately at the end
𝑑2 𝑣𝑏
of its linearly elastic region and the curvature is = 6,982 ⋅ 10−4 𝑚𝑚−1 , making the value of
𝑑𝑥 2
𝑏 2 𝑑2 𝑣𝑏
equal to 5,28 which makes the stiffening effect 1,0351 if the poissons ratio is taken as 0,3
𝑡 𝑑𝑥 2
and the value of the stiffening effect is interpolated from the values in Table 2.
𝑑2 𝑣𝑏
Similarly, for a foam plate at a load of 50 Newton, the curvature is = 6,146 ⋅ 10−6 𝑚𝑚−1 and
𝑑𝑥 2
𝑏 2 𝑑2 𝑣𝑏
the corresponding value of equal to 0,0465 𝑚𝑚−1 , which makes the stiffening effect
𝑡 𝑑𝑥 2
approximately 1 regardless of the poissons ratio.
It should be noted that these values of the stiffening effect is calculated from the moment at the
center of the beam and thus the stiffening effect of preventing anticlastic curvature reduces linearly
to 1 at the supports so for the overall stiffness of the beam, the stiffening effect is even lower.
Within this project, it is not attempted to reproduce [8] for laminated plates to evaluate the
assumption that 3-point bending of specimen for formula student can be evaluated as beams.
Instead, beam models for laminated beams are produced and compared to experimental results.

Page 16 af 125
Figure 5: Anticlastic curvature seen at edges of bent foam plate

3.6 Bernoulli-Euler beams (Classical beam theory)


Profiles that are much more slender and thin than they are long can be modeled as beams whose
displacement and stress fields can be described from only one spatial coordinate. This one-
dimensionality of beams is an abstraction from the fact that solids are three-dimensional. The
simplification to one-dimensionality comes from the application of an assumed deformation field to
a rectangular region of plane stress. [6]
This abstraction can be modeled through numerous different theories of which the Bernoulli-Euler
and Timoshenko beam theories are the most common as these are among the simplest models and
have been in use for practically all of modern mechanical engineering. The displacement field that
is assumed for a Bernoulli-Euler beam is that the transverse deflection is constant through the
thickness of the beam and that the displacement in the axial direction varies at most linearly with
respect to the transverse slope of the centerline as given in (3-41).
𝑑𝑣0 (3-41)
𝑢(𝑥, 𝑦) = 𝑢0 (𝑥) − 𝑦
𝑑𝑥
𝑣(𝑥, 𝑦) = 𝑣0 (𝑥)
The strains are calculated by elasticity theory in (3-42), where it is seen that due to the assumed
displacement field, the through-thickness shear is zero and thus cross-sections through the beam
remain plane and normal to the centerline after deformation. Since shearing forces can exist in

Page 17 af 125
Bernoulli-Euler beams, this theory will overestimate the stiffness as the shearing deformation is
neglected.
𝑑𝑢 𝑑𝑢0 𝑑 2 𝑣0 (3-42)
𝜀1 = = −𝑦
𝑑𝑥 𝑑𝑥 𝑑𝑥 2
𝑑𝑣
𝜀2 = =0
𝑑𝑦
𝑑𝑢 𝑑𝑣 𝑑𝑣0 𝑑𝑣0
𝜀6 = + =− + =0
𝑑𝑦 𝑑𝑥 𝑑𝑥 𝑑𝑥
By rule of thumb, the Bernoulli-Euler beam theory can be used for homogeneous, isotropic beams
which are ten times as long as they are thick as the shear deformation becomes small compared to
the deformation caused by the bending moment. For anisotropic or laminated beams, this rule of
thumb does not apply as the shearing deformation can be of much larger magnitude and especially
so when the beam is laminated as a sandwich with a flexible core. For a homogeneous beam, the
deflection curve can be solved from the differential equation in (3-43), which is found by
integrating the stresses over the cross-section where the coordinate system coincides with the
geometrical center of the cross-section.
𝑑 2 𝑣0 (3-43)
𝐸𝐼 ⋅ = 𝑀(𝑥)
𝑑𝑥 2
For a laminate in which 𝑀(𝑥) = 𝑀𝑥𝑥 ⋅ 𝑏 is the only nonzero moment, equation (3-44) provides an
analog for the bending stiffness, as laminate theory gives this as the inverse of the factor that
couples the bending moment to the curvature.
𝑏 (3-44)
(𝐸𝐼)𝑐 =
̅11
𝐷
The Bernoulli-Euler beam exists as a special case of the Timoshenko beam in which the shear
stiffness is infinite and thus the stiffness matrix for the finite element can be taken from equation
(3-74) by setting Φ = 0. Doing this, the degree of freedom corresponding to the rotation of the
cross-section equals the slope of the beam centerline and the element is said to have 𝐶 2 -continuity
which means that the slope of the centerline is continuous across elements, even if they have
different stiffnesses or if the node has a load on it.

3.7 Timoshenko beams (First-order shear deformation theory)


The Timoshenko beam theory predicts the deflection of a beam more accurately than the Bernoulli-
Euler theory by assuming the shear deformation to be of first order through the thickness of the
beam. This assumption makes the shear strain constant through the thickness rather than the
parabolic shape that elasticity theory predicts to make the shear stress zero at the surfaces of the
beam. To account for this physically incorrect assumption of shear stress, the Timoshenko beam
theory employs a shear correction factor, 𝜅 as described later in this chapter. The displacement field
in (3-45) which is assumed for the Timoshenko beam is similar to that of the Bernoulli-Euler beam
except that the axial displacement varies linearly with respect to the rotation of the cross-section
which may be different than the slope of the centerline.

Page 18 af 125
𝑢(𝑥, 𝑦) = 𝑢0 (𝑥) − 𝑦𝜙𝑥 (3-45)

𝑣(𝑥, 𝑦) = 𝑣0 (𝑥)
This implies that cross-sections remain plane but are not restricted to be normal to the centerline
and as shown in equation (3-46)(3-47), the shear strain does not vary through the thickness of the
beam.
𝑑𝑢 𝑑𝑢0 𝑑𝜙𝑥 (3-46)
𝜀1 = = −𝑦
𝑑𝑥 𝑑𝑥 𝑑𝑥
𝑑𝑣
𝜀2 = =0
𝑑𝑦
𝑑𝑢 𝑑𝑣 𝑑𝑣0
𝜀6 = + = − 𝜙𝑥
𝑑𝑦 𝑑𝑥 𝑑𝑥
The rotation of the cross-section is shown in Figure 6, where it can be seen that the slope of the
centerline is greater than the rotation of the cross-section as the difference between these is the
shearing strain. The rotation of the cross-section can be interpreted as the part of the centerline
slope that is caused by bending deformations as shearing deformations will cause a slope of the
centerline without any rotation of the cross-section.

Figure 6: Timoshenko beam cross-section rotation and centerline slope

The development of a shear deformable finite element can be done by shape functions but unless
the chosen shape functions describe the deformation exactly, shear stiffening effects can occur
which requires reduced integration to get rid of. [9].
To develop an exact finite element for laminated beams in a way that is likely to be understood by
the students who will be using the element, it is derived using the direct stiffness method similarly
to how the element for homogeneous beams are developed in [4].

Page 19 af 125
When only an axial force and a bending moment is allowed to act upon the beam, the axial strain is
given by equation (3-47) which is then integrated once to yield the displacement field by equation
(3-48).
̅
𝑑𝑢 𝐴11 𝐵̅11 (3-47)
= 𝑁(𝑥) − 𝑀(𝑥)
𝑑𝑥 𝑏 𝑏
𝐴̅11 𝐵̅11 (3-48)
𝑢(𝑥) = ∫ 𝑁(𝑥)𝑑𝑥 − ∫ 𝑀(𝑥)𝑑𝑥
𝑏 𝑏

Where 𝑁(𝑥) = 𝑏𝑁𝑥𝑥 is an axial tensile load and 𝑀(𝑥) = 𝑏𝑀𝑥𝑥 is a bending moment.
The transverse displacement in equation (3-49) is split into two components, such that one is the
contribution from axial strain due to bending and the other is the contribution from shearing strain.
𝑣 = 𝑣𝑏 + 𝑣𝑠 (3-49)

The displacements due to shearing strain is assumed such that the derivative of the displacement
equals the shearing strain as shown in equation (3-50) while the bending curvature is defined from
laminate theory and are given in equation (3-51). By integrating these and adding them, the
transverse deflection of the beam can be described by equation (3-52).
𝑑𝑣𝑠 𝑉(𝑥) (3-50)
=−
𝑑𝑥 𝑏𝜅𝐴55
𝑑2 𝑣𝑏 𝐵̅11 ̅11
𝐷 (3-51)
= − 𝑁(𝑥) + 𝑀(𝑥)
𝑑𝑥 2 𝑏 𝑏
𝐵̅11 ̅11
𝐷 ∫ 𝑉(𝑥)𝑑𝑥 (3-52)
𝑣(𝑥) = − ∫ ∫ 𝑁(𝑥)𝑑𝑥 𝑑𝑥 + ∫ ∫ 𝑀(𝑥)𝑑𝑥 𝑑𝑥 −
𝑏 𝑏 𝑏𝜅𝐴66

The beam element is loaded as shown on Figure 7 and by taking a cross-section at an arbitrary
position, it can be shown that the loads in equations (3-47) to (3-52) can be defined from the end
loadings as equation (3-53).
𝑁(𝑥) = −𝐹1 , 𝑉(𝑥) = 𝐹2 , 𝑀(𝑥) = 𝐹2 𝑥 − 𝐹3 (3-53)

Page 20 af 125
Figure 7: Boundary conditions for timoshenko beam with arbitrary axial displacement at one end.

Under these loads, the axial displacement, the transverse displacement and the derivative of the
transverse displacement become as shown in equations (3-54), (3-55) and (3-56) respectively. The
components of the stiffness matrix can then be derived by application of boundary conditions
followed by solving for the values of each of the degrees of freedom at the nodes and dividing these
by the loads. To show the process, this is shown for the boundary condition in Figure 7 where the
only non-zero degree of freedom is an arbitrary displacement at the leftmost end of the beam.
̅
𝐴11 𝐵̅11 𝐹2 𝑥 2 (3-54)
𝑢(𝑥) = − 𝐹1 𝑥 − ( − 𝐹3 𝑥) + 𝐶1
𝑏 𝑏 2

𝐵̅11 −𝐹1 𝑥 2 𝐷̅11 𝐹2 𝑥 3 𝐹3 𝑥 2 𝐹2 𝑥 (3-55)


𝑣(𝑥) = − + ( − )− + 𝐶2 𝑥 + 𝐶3
𝑏 2 𝑏 6 2 𝑏𝜅𝐴66

𝑑𝑣 𝐵̅11 ̅11 𝐹2 𝑥 2
𝐷 𝐹2 (3-56)
= 𝐹1 𝑥 + ( − 𝐹3 𝑥) − + 𝐶2
𝑑𝑥 𝑏 𝑏 2 𝑏𝜅𝐴66

By application of the boundary conditions at 𝑥 = 0, the constants of integration are solved for.
𝑢(0) = 𝐶1 = 𝑢1 (3-57)

𝑣(0) = 𝐶3 = 0 (3-58)

𝑑𝑣 𝐹2 𝐹2 (3-59)
(0) = − + 𝐶2 = − ⇒ 𝐶2 = 0
𝑑𝑥 𝑏𝜅𝐴55 𝑏𝜅𝐴55

𝑑𝑢
Worth noting here is that the built-in boundary condition is taken as 𝑑𝑧 = 0 at 𝑦 = 0, which implies
𝑑𝑣𝑏 𝑑𝑣 𝑑𝑣𝑠 𝐹
𝜙𝑥 = = 0 as the cross-section is prevented from rotating and thus 𝑑𝑥 = = − 𝑏𝜅𝐴2 . From the
𝑑𝑥 𝑑𝑥 66
boundary conditions at 𝑥 = 𝑙𝑒 , the forces and finally an expression for the arbitrary displacement
are derived in equations (3-60) through (3-65).

Page 21 af 125
𝑑𝑣 𝐵̅11 ̅11 𝐹2 𝑙𝑒2
𝐷 𝐹2 𝐹2 (3-60)
(𝑙𝑒 ) = 𝐹1 𝑙𝑒 + ( − 𝐹3 𝑙𝑒 ) − =−
𝑑𝑥 𝑏 𝑏 2 𝑏𝜅𝐴66 𝑏𝜅𝐴66

𝐵̅11 𝐹2 𝑙𝑒 (3-61)
⇒ 𝐹3 = 𝐹1 +
𝐷̅11 2
𝐵̅11 𝐹1 𝑙𝑒2 𝐷̅11 𝐹2 𝑙𝑒3 𝐹3 𝑙𝑒2 𝐹2 𝑙𝑒 (3-62)
𝑣(𝑙𝑒 ) = + ( − )− =0
𝑏 2 𝑏 6 2 𝑏𝜅𝐴66

⇒ 𝐹2 = 0 (3-63)

𝐴̅11 𝐵̅11 𝐹2 𝑙𝑒2 (3-64)


𝑢(𝑙𝑒 ) = − 𝐹𝑙 − ( − 𝐹3 𝑙𝑒 ) + 𝑢1 = 0
𝑏 1𝑒 𝑏 2

𝐴̅11 (𝐵̅11 )2 (3-65)


⇒ 𝑢1 = 𝐹1 𝑙𝑒 − 𝐹𝑙
𝑏 𝑏𝐷̅11 1 𝑒

The forces and moment at the other end of the beam are expressed in terms of 𝐹1 , 𝐹2 and 𝐹3 , derived
from static equilibrium of the beam and thus the column of the stiffness matrix relating the forces to
𝑢1 can be found similarly to how the first coefficient is shown in equation (3-66).
−1 (3-66)
̅ (𝐵̅11 )2
𝑏 (𝐴11 − ̅ )
𝐹1 𝐷11
𝑘11 = =
𝑢1 𝑙𝑒
By switching which end is allowed to have an arbitrary axial displacement, the coefficients relating
to 𝑢2 are found. Similarly, the remaining coefficients of the stiffness matrix are found by applying
boundary conditions of an arbitrary transverse displacement or an arbitrary rotation of the cross-
section at one end, as shown in Figure 8 and Figure 9.

Figure 8: Beam element with an arbitrary transverse Figure 9: Beam element with an arbitrary rotation of the cross-
displacement at one end sectionat one end

To simplify the equations and keep the contribution from shear deformation clear, the constant Φ is
12𝐸𝐼
introduced, analog to the constant equal to 𝐺𝜅𝐴𝑙2 for isotropic beams which is presented in [4].
𝑒

̅11
12𝑏𝐷 −1 (3-67)
Φ=
𝑏𝜅𝐴66 𝑙𝑒2
All of the coefficients of the stiffness matrix in equation (3-74) is given in equations (3-68) through
(3-73).

Page 22 af 125
−1 (3-68)
̅ − (𝐵̅11 )2
𝑏 (𝐴11 ̅11 )
𝐷
𝑘11 = 𝑘44 = −𝑘14 =
𝑙𝑒
𝑏𝐵̅11 (3-69)
𝑘13 = 𝑘46 = −𝑘16 = −𝑘34 =
̅ 𝐷
𝑙𝑒 (𝐴11 ̅11 − (𝐵̅11 )2 )

12𝑏𝐷 ̅11
−1 (3-70)
𝑘22 = 𝑘55 = −𝑘25 =
𝑙𝑒3 (1 + Φ)
6𝑏𝐷̅11
−1 (3-71)
𝑘23 = 𝑘26 = −𝑘35 = −𝑘56 =
(1 + Φ)𝑙𝑒2
̅ 𝐷
𝐴11 ̅11 (4 + Φ) − 3(𝐵̅11 )2 (3-72)
𝑘33 = 𝑘66 =𝑏
̅11 (1 + Φ)𝑙𝑒 (𝐷
𝐷 ̅11 𝐴11
̅ − (𝐵̅11 )2 )
̅ 𝐷
𝑏(𝐴11 ̅11 (2 − Φ) − 3(𝐵̅11 )2 ) (3-73)
𝑘36 =
̅11 (1 + Φ)(𝐷
𝑙𝑒 𝐷 ̅11 𝐴11
̅ − (𝐵̅11 )2 )

𝑘11 0 𝑘13 𝑘14 0 𝑘16 𝑢1 𝐹1 (3-74)


0 𝑘22 𝑘23 0 𝑘25 𝑘26 𝑣1 𝐹2
𝑘13 𝑘23 𝑘33 𝑘34 𝑘35 𝑘36 𝜙1 𝐹3
=
𝑘14 0 𝑘34 𝑘44 0 𝑘46 𝑢2 𝐹4
0 𝑘25 𝑘35 0 𝑘55 𝑘56 𝑣2 𝐹5
[𝑘16 𝑘26 𝑘36 𝑘46 𝑘56 𝑘66 ] {𝜙2 } {𝐹6 }

Because of the coupling between extension and bending, an element without extensional DoF
cannot be constructed simply by omitting rows and columns of the stiffness matrix as this would be
equivalent to locking the axial displacement of all nodes, thus creating axial loads due to bending
which has an effect on the bending stiffness. Instead, the development of the element is done while
setting 𝐹1 = 𝐹4 = 0 and only solving in the equations for the displacement and its derivative. The
̅11
stiffness matrix then takes a form similar to that in [2] and [4], just with 𝐸𝐼 replaced by 𝑏𝐷 −1
and
with 𝐴𝐺 replaced by 𝑏𝐴66 . The use of 𝐴66 rather than 𝐴55 is due to the beam having the y-direction
as the through-thickness direction.
12 6𝑙𝑒 −12 6𝑙𝑒 𝑣1 𝐹2 (3-75)
̅11
𝑏𝐷 −1
6𝑙𝑒 𝑙𝑒2 (4+ Φ) −6𝑙𝑒 𝑙𝑒2 (2
− Φ) 𝜙 1 𝐹
{ 𝑣 } = { 3}
3
𝐿𝑒 (1 + Φ) −12 −6𝑙𝑒 12 −6𝑙𝑒 2 𝐹5
[ 6𝑙𝑒 2
𝑙𝑒 (2 − Φ) −6𝑙𝑒 6𝑙𝑒 ] 𝜙2 𝐹6

Page 23 af 125
3.7.1 Shear correction factor
The Timoshenko beam theory employs a physically incorrect assumption of the shear strain and
makes use of a shear correction factor which reduces the shear stiffness to the same value as a beam
with a physically correct shear stress distribution.
The shear stress correction factor is calculated as described in [5], by taking the ratio of the strain
energy of the assumed shear stress distribution to the strain energy of the actual shear stress
distribution. For a homogeneous beam, the shear stress given by elasticity theory is 𝜎6𝑐 =
3𝑉𝑥 2𝑦 2 𝑉
(1 − ( ) ) while the shear stress from first order shear deformation theory is 𝜎6𝑓 = 𝑥 . The
2𝑏ℎ ℎ 𝑏ℎ
shear strain energies are then as in
3𝑉𝑥2 (3-76)
𝑈𝑠𝑐=
5𝐺𝑏ℎ
𝑓 𝑉𝑥2 (3-77)
𝑈𝑠 =
2𝜅𝐺𝑏ℎ

5
So the shear correction factor for the rectangular cross-section homogeneous beam are 𝜅 = 6. The
actual shear stress distribution for a laminated beam is not equal to the shear stress distribution of a
homogeneous beam, so the commonly used value is incorrect.
The axial stress caused by bending and stretching the laminate can be used to calculate the actual
shear stress by solving for it in equation (3-78) which forms part of the 2D elastic equilibrium
equations. Once the actual shear stress through the lamina is calculated, the potential shear strain
energy can be calculated and set equal to the potential energy of the assumed strain distribution so
that the shear correction factor can be solved for.
𝜕𝜎1 𝜕𝜎6 (3-78)
+ =0
𝜕𝑥 𝜕𝑦
𝜕𝜎6 𝜕𝜎2 (3-79)
+ =0
𝜕𝑥 𝜕𝑦

Using this method, the algebraic expression for the shear stress correction factor in equation (3-80)
is developed by [11] but is here given with the y being the through-thickness direction. This shear
stress correction factor is calculated from the assumption that the bending moment whose derivative
equals the shear force are the only loads. Due to the complexity of the expression in (3-80),
attempts at adapting it to include normal forces has not been successful in this project but since the
coupling between axial force and transverse deflection of an unsymmetric beam are due to bending,
axial loads are assumed not to have any coupling to shear deformation.

Page 24 af 125
−1 (3-80)
(𝐴56 )2
𝜅 = (𝐴55 − )
𝐴66

𝑁
1
⋅ ∑ 2
(𝑘)
𝑘=1 (𝑘)
(𝑄̅56 )
𝑄̅55 − (𝑘)
( ( 𝑄̅66 )
−1

𝑅𝑘 2 𝑉𝑘
𝑃𝑘 (𝑦𝑘 − 𝑦𝑘−1 ) + 2
(𝑦𝑘 − 𝑦𝑘−41 ) + (𝑦𝑘3 − 𝑦𝑘−1
3
)
⋅( 2 3 )
𝑊𝑘 4 4
𝑋𝑘 5 5
+ (𝑦 − 𝑦𝑘−1 ) + (𝑦𝑘 − 𝑦𝑘−1 )
4 𝑘 5
)

Where
𝐽𝑘2 𝑦𝑘4
𝑃𝑘 = 𝑇𝑘2 + 𝐻𝑘2 𝑦𝑘2 − 2𝑇𝑘 𝐻𝑘 𝑦𝑘 + 𝑈𝑘2 + − 𝑈𝑘 𝐽𝑘 𝑦𝑘2 + 2𝑇𝑘 𝑈𝑘 − 𝑇𝑘 𝐽𝑘 𝑦𝑘2 − 2𝐻𝑘 𝑈𝑘 𝑦𝑘 + 𝐻𝑘 𝐽𝑘 𝑦𝑘3
4
𝑅𝑘 = 2𝑇𝑘 𝐻𝑘 − 2𝐻𝑘2 𝑦𝑘 + 2𝐻𝑘 𝑈𝑘 − 𝐻𝑘 𝐽𝑘 𝑦𝑘2
𝐽𝑘2 𝑦𝑘2
𝑉𝑘 = 𝐻𝑘2 − + 𝑈𝑘 𝐽𝑘 + 𝑇𝑘 𝐽𝑘 − 𝑦𝑘 𝐻𝑘 𝐽𝑘
2
𝑊𝑘 = 𝐻𝑘 𝐽𝑘
𝐽𝑘2
𝑋𝑘 =
4
𝑘−1

𝑇𝑘 = ∑ 𝐻𝑚 (𝑦𝑚 − 𝑦𝑚−1 )
𝑚=1
𝑘−1
𝐽𝑚 2 2
𝑈𝑘 = ∑ (𝑦 − 𝑦𝑚−1 )
2 𝑚
𝑚=1

(𝑘)
𝐻𝑘 = ∑ 𝑄̅1𝑖 𝐵̅1𝑖
𝑖=1,2,6

(𝑘) ̅
𝐽𝑘 = ∑ 𝑄̅1𝑖 𝐷 1𝑖
𝑖=1,2,6

By use of the correction factor, the stiffness of the Timoshenko beam equals that of a beam with the
correct shear stress distribution, but it does not correct the fact that the boundary conditions
influence the stiffness. The concentrated load at the center of the beam in 3-point bending will mean
that the slope of the shear deformation has opposite sign on either side of the load and thus even

Page 25 af 125
though the slope of the bending deflection is continuous, the slope of the centerline is not
continuous and thus the beam is said to have only 𝐶 0 -continuity because none of the derivatives are
continuous. As shown on Figure 10, the calculated shape of the beams has a sharp bend which
would not be physically possible without fracturing the composite fibers. Another consequence of
the assumed boundary conditions and degree of continuity is that any beam that overhang the
supports have no stresses and thus no influence on the calculated stiffness. The influence of slightly
incorrect boundary conditions is argued by Timoshenko to only be of little influence to the overall
stiffness due to Saint-Venant’s principle. [6] This principle states that small changes to the
distribution of a load in a way that is statically equivalent only produces localized changes to the
stress distribution.

Figure 10: Sketch of Timoshenko beam in 3 point bending

Further discussion of the boundary conditions exist in section 3.8.2 where the Timoshenko beam
model is compared to another beam model for which the only difference in the stiffness of the beam
element exist in the boundary conditions.
3.7.2 Stress calculations
In order to calculate the stresses of a finite element, the strain displacement matrix derived from
shape functions are usually used but since shape functions have not been used to derive the beam
elements in this project, it is necessary to calculate them now in order to calculate the stresses.
The shape functions are found from the equations used to derive the stiffness matrix by diving the
function for the displacements and rotation by the degrees of freedom at the nodes. When the values
of the degrees of freedom at the nodes are calculated by the stiffness matrix and external loads,
these can be multiplied by the shape functions 𝑁𝑖𝑗 to yield the values of the degrees of freedom
within the element as shown in equation (3-81) and (3-95). The shape functions are given in Table
3.

Page 26 af 125
𝑢1 (3-81)
𝑣1
𝒩11 𝒩12 𝒩13 𝒩14 𝒩15 𝒩16 𝑢(𝑥)
𝜙1
[𝒩21 𝒩22 𝒩23 𝒩24 𝒩25 𝒩26 ] 𝑢 = { 𝑣(𝑥) }
2
𝒩31 𝒩32 𝒩33 𝒩34 𝒩35 𝒩36 𝑣 𝜙(𝑥)
2
{𝜙2 }

𝒖(𝒙) 𝒗(𝒙) 𝝓(𝒙)


𝒖𝟏 𝑥 𝒩21 = 0 𝒩31 = 0
𝒩11 = (1 − )
𝑙𝑒
𝒗𝟏 𝐵̅ (𝑙𝑒 + 2𝑥)(−𝑥 + 𝑙𝑒 )2 6𝑥(−𝑥 + 𝑙𝑒 )
6 ̅11 (−𝑥 + 𝑙𝑒 ) 𝒩22 = 𝒩32 = −
𝐷11 𝑙𝑒2 (1 + Φ) 𝑙𝑒3 (1 + Φ)
𝒩12 =
𝑙𝑒3 (1 + Φ)
𝝓𝟏 𝐵̅ (𝑙𝑒 − 𝑥)2 (Φ𝑙𝑒 − 2𝑥) 𝒩33 =
3 ̅11 (𝑙𝑒 − 𝑥)𝑥 𝒩23 = −
𝐷11 2𝑙𝑒2 (1 + Φ)
𝒩13 = (𝑙𝑒 − 𝑥)((1 + Φ)𝑙𝑒 − 3𝑥)
𝑙𝑒2 (1 + Φ)
𝑙𝑒2 (1 + Φ)
𝒖𝟐 𝑥 𝒩24 = 0 𝒩34 = 0
𝒩14 =
𝑙𝑒
𝒗𝟐 𝐵̅ 3𝑥 2 2𝑥 3 6𝑥(−𝑥 + 𝑙𝑒 )
−6 ̅11 (−𝑥 + 𝑙𝑒 ) (− 3 )+Φ 𝒩35 =
𝐷11 𝑙𝑒2 𝑙𝑒 𝑙𝑒3 (1 + Φ)
𝒩15 = 𝒩25 =
𝑙 3 (1 + Φ) (1 + Φ)

𝝓𝟐 𝐵̅ 𝒩26 = (𝜙 − 2)𝑙𝑒 𝑥 + 3𝑥 2
3 ̅11 (𝑙𝑒 − 𝑥)𝑥 𝒩36 =
𝐷11 𝑙𝑒2 (1 + 𝜙)
𝒩16 = −𝑙𝑒3 𝜙 + 𝑥 2 𝑙𝑒 (Φ − 2) + 2𝑥 3
(1 + Φ)𝑙𝑒2
2𝑙𝑒2 (1 + 𝜙)
Table 3: Shape functions for timoshenko beam element

To find the axial strain and the bending curvature, only the shape functions related to the axial and
rotational DoF, in equation (3-82) and (3-83) are used.
⌊𝒩1 ⌋ = ⌊𝒩11 𝒩12 𝒩13 𝒩14 𝒩15 𝒩16 ⌋ (3-82)

⌊𝒩3 ⌋ = ⌊𝒩31 𝒩32 𝒩33 𝒩34 𝒩35 𝒩36 ⌋ (3-83)

These are then differentiated with respect to x to gain the strain-displacement functions in equation
(3-84) and (3-85) with the coefficients given in table
⌊𝔅1 ⌋ = ⌊𝔅11 𝔅12 𝔅13 𝔅14 𝔅15 𝔅16 ⌋ (3-84)

⌊𝔅3 ⌋ = ⌊𝔅31 𝔅32 𝔅33 𝔅34 𝔅35 𝔅36 ⌋ (3-85)

Page 27 af 125
𝒅𝒖 𝒅𝝓
𝒅𝒙 𝒅𝒙
𝒖𝟏 1 𝔅31 = 0
𝔅11 = −
𝑙𝑒
𝒗𝟏 𝐵̅ 12𝑥 − 6𝑙𝑒
6 ̅11 (−2𝑥 + 𝑙𝑒 ) 𝔅32 =
𝐷11 𝑙𝑒3 (1 + Φ)
𝔅12 =
𝑙𝑒3 (1 + Φ)
𝝓𝟏 𝐵̅ (−Φ − 4)𝑙𝑒 + 6𝑥
3 ̅11 (𝑙𝑒 − 2𝑥) 𝔅33 =
𝐷11 𝔅12 (1 + Φ)𝑙𝑒2
𝔅13 = 2
=
𝑙𝑒 (1 + Φ) 2
𝒖𝟐 1 𝔅34 = 0
𝔅14 =
𝑙𝑒

𝒗𝟐 𝐵̅ −12𝑥 + 6𝑙𝑒
6 ̅11 (−2𝑥 + 𝑙𝑒 ) 𝔅35 = = −𝔅32
𝐷11 𝑙𝑒3 (1 + Φ)
𝔅15 =− = −𝔅12
𝑙𝑒3 (1 + Φ)

𝝓𝟐 𝐵̅ (Φ − 2)𝑙𝑒 + 6𝑥
3 ̅11 (𝑙𝑒 − 2𝑥) 𝔅36 =
𝐷11 𝔅12 (1 + Φ)𝑙𝑒2
𝔅16 = 2
=
𝑙𝑒 (1 + Φ) 2
Table 4: Strain-displacement functions for laminated Timoshenko beam element

It is worth noting that as a special case, the strain-displacement functions in equation (3-98) reduce
to those of a bar element for a symmetric laminate or homogeneous beam where 𝐵̅11 = 0 and that
those in (3-85) reduce to those of a Bernoulli-Euler beam when Φ = 0. With these strain-
displacement functions, the axial strains are calculated per equation (3-86).
(
𝜀1 0) ⌊𝔅1 ⌋ (3-86)
{ }=[ ] {𝑑}
(1)
𝜀1 ⌊𝔅3 ⌋

To find the in-plane stresses the beam is now assumed to be a plate strip under these strains along
with other to-be-determined in-plane strains. A bending moment and an in-plane force only in the x-
direction are assumed to be the only external loads. The axial force and the bending moment can be
calculated from (3-35) but as the transverse and in-plane shear strains aren’t explicitly calculated in
the beam calculations, calculating by (3-87) and (3-88) are incorrect for this element. For the beam
element where the axial degree of freedom is omitted, calculating the bending moment from (3-88)
̅11
gives the correct result by omitting the terms relating to midplane strain and using 𝐷 −1
rather than
the stiffnesses 𝐷11 , while 𝑁𝑥 = 0.
𝑑𝑢 𝑑𝜙 (3-87)
𝑁𝑥 = 𝑏𝐴11 − 𝑏𝐵11 = 𝑏(𝐴11 ⌊𝔅1 ⌋ − 𝐵11 ⌊𝔅3 ⌋){𝑑}
𝑑𝑥 𝑑𝑥

Page 28 af 125
𝑑𝑢 𝑑𝜙 (3-88)
𝑀𝑥 = −𝑏𝐵11 + 𝑏𝐷11 = 𝑏(−𝐵11 ⌊𝔅1 ⌋ + 𝐷11 ⌊𝔅3 ⌋){𝑑}
𝑑𝑥 𝑑𝑥
For the beam element with axial displacement where (3-87) and (3-88) are incorrect, the external
loads on an element are instead calculated by multiplying the element stiffness matrix in (3-74) by
the nodal displacements and then calculating the in-plane force and bending moment from (3-53).
Using the calculated force and moment, the in-plane strains and bending curvatures can now be
calculated.
(0) (3-89)
𝜀1 ̅
𝐴11 𝐵̅11
(0) [𝐴̅] 𝑁𝑥 [𝐵̅ ] 𝑀𝑥 𝑁𝑥 𝑀𝑥
{𝜀3 } = { 0}− { 0 } = {𝐴̅31 } ⋅ − {𝐵̅31 } ⋅
𝑏 𝑏 𝑏 𝑏
𝜀
(0) 0 0 𝐴̅41 𝐵̅41
4
(1) (3-90)
𝜀1 𝐵̅11 ̅11
𝐷
[𝐵̅ ]𝑇 𝑁𝑥 ̅ ] 𝑀𝑥
[𝐷 𝑁𝑥 𝑀
(1)
{𝜀3 } = − {0}+ { 0 } = − {𝐵̅13 } ⋅ ̅31 } ⋅ 𝑥
+ {𝐷
𝑏 𝑏 𝑏 𝑏
𝜀
(1) 0 0 𝐵̅14 ̅41
𝐷
4

Thus, the in-plane stresses in the 𝑖 ′ 𝑡ℎ lamina at a distance 𝑧𝑖 from the midplane, in the global
coordinate system are calculated per (3-91).
(0) (1) (3-91)
𝜎̅1 𝜀1 𝜀1
{𝜎̅3 } = [𝑄̅ ]𝑖 ⋅ ({𝜀3(0) } − 𝑧 {𝜀3(1) })
𝜎̅4 𝑖 (0) (1)
𝜀4 𝜀4
This stress is then transformed into the coordinate system of the lamina by a transformation matrix
[𝑇], which is similar to the one in (3-22) but rearranged so that it rotates around the y-axis instead
of the z-axis.
𝜎1 𝜎̅1 (3-92)
𝜎
{ 3} = [𝑇] −1 𝜎
{ ̅3 }
𝜎6 𝑖 𝜎̅6 𝑖

With the lamina stresses in the plane of the coordinate system, plane failure criteria can be applied.
While the through-thickness shear stress can be found directly from the equations of the
Timoshenko beam, it is not very useful, as it would be calculated to be a constant value and for a
Bernoulli-Euler beam, it cannot be calculated from the beam theory at all. Similarly, through-
thickness compressive stresses are not calculated by any of these beam models. Instead, the
through-thickness stresses may be computed from the equilibrium equations of static 3D elasticity
in equation (3-93).
𝜕𝜎1 𝜕𝜎6 𝜕𝜎4 (3-93)
+ + =0
𝜕𝑥 𝜕𝑦 𝜕𝑧

Page 29 af 125
𝜕𝜎6 𝜕𝜎2 𝜕𝜎5
+ + =0
𝜕𝑥 𝜕𝑦 𝜕𝑧
𝜕𝜎4 𝜕𝜎5 𝜕𝜎3
+ + =0
𝜕𝑥 𝜕𝑦 𝜕𝑧
Due to the beam assumption, the derivatives with respect to the transverse direction are zero and
thus (3-93) reduces to the plane stress equilibrium equations in (3-78) and (3-79). When the bending
moment and shear force are the only loads, the transverse normal stress is a multiplication of the
derivative of the shear force which is zero in these beam elements as all loads are applied as point
loads at the nodes.
The transverse shear stresses at any point through the thickness is calculated in the k’th layer as in
equation (3-94) by integrating up until the coordinate, y, at which the stress is wanted known. The
integration constant 𝐺 (𝑘) equals the stress at the bottom surface of the k’th layer at 𝑦𝑘−1 .
𝑦 (3-94)
(𝑘) 𝑉𝑥 (𝑘)
𝜎6 = −∫ ( ∑ 𝑄̅1𝑖 (−𝐵̅1𝑖 + 𝑦𝐷
̅1𝑖 )) 𝑑𝑦
ℎ 𝑏
2 𝑖=1,3,4

𝑉𝑥 (𝑘) (𝑘) (𝑘)


=− (−(𝑄̅11 (𝐵1 )11 + 𝑄̅12 (𝐵1 )12 + 𝑄̅16 (𝐵1 )16 )(𝑦 − 𝑦𝑘−1 )
𝑏
2 )
(𝑦 2 − 𝑦𝑘−1
̅ (𝑘) ̅ (𝑘) ̅ (𝑘)
+ (𝑄11 (𝐷1 )11 + 𝑄12 (𝐷1 )12 + 𝑄16 (𝐷1 )16 ) ( ))
2
+ 𝐺 (𝑘)

On Figure 11, the shear stresses of an asymmetric beam with 14 layers in the top skin, 4 layers in
the bottom skin and a 25mm thick core are shown. The stresses in Figure 11 are calculated at
10.000 equidistant points and vary parabolically through each layer even though it doesn’t look like
it. The apparent linear variation of shear stress through each layer are due to the laminated structure
with large differences in shear stiffness.

Page 30 af 125
Figure 11: Through-thickness shear stress of an asymmetric beam

3.8 Other beam theories


Many other beam theories have been developed to account for the shear deformation in beams.
These theories all apply some assumptions to the shear strain through the thickness of the beam
where some assume the shear deformation to be of higher order to fulfil the boundary condition of
vanishing shear stresses at the surfaces of the beam and others assume certain features to be
negligible in order to simplify the calculations. A few of these other theories are described in this
section and while these are not chosen for the modeling of 3 point bending in this project, it serves
to show that there are other possible methods.
3.8.1 Sandwich theory
Among the simpler theories are sandwich theory, where the face sheets are assumed thin and stiff
relative to the core. [12] In this theory, the shear stresses in the face sheets are assumed negligible,
leading to all the shear deformation being due to a constant shear through the core. The shear
deformation of a beam at distance 𝑥 from the left end are then calculated per equation (3-95)(3-80),
in which 𝑉 are the shear force, 𝑏 are the width of the beam, 𝑡𝑐 are the thickness of the core and 𝐺𝑐
are the shear modulus of the core.
𝑉⋅𝑥 (3-95)
𝑣𝑠 (𝑥) =
𝑏𝑡𝑐 𝐺𝑐

Page 31 af 125
The shear deflection is then added to the deflection due to bending. When the face sheets are very
thin or if they are not sufficiently stiff in shear, this theory will overestimate the stiffness as the
shear stress tends towards the distribution of a homogeneous beam. On the other hand, when the
face sheets are thick, the stiffness is underestimated due to not including the potential shear strain
energy in the face sheets. On Figure 12, the shear deflection under a unit load is compared to the
solution of a Timoshenko beam as derived in section 3.7 with a shear correction factor as described
in section 3.7.1. The beam is assumed to have a 25mm thick core of M80 foam and face sheets
consisting of Gurit SE84-LV HEC prepreg with the fibers in 0-degree direction.

Figure 12: Comparison of shear stiffness of sandwich and timoshenko theories

When doing a similar comparison where the face sheets are composed of alternating 0 degree and
90 degree layers, the error rises more slowly than when it consists of only 0 degree layers but is still
significant with an error of 41.2% when the face sheets have a thickness of 4,256mm with 14 layers.
With these materials, the shear stiffness will be underestimated for all practical thicknesses of face
sheets but since the both the sign and magnitude of the error depends on the material stiffnesses and
thicknesses, the sandwich theory cannot be thought of as a generally applicable theory to predict the
stiffness of a laminated beam.

3.8.2 R. P. Shimpi single-variable shear deformation theory


This theory presented by Shimpi applies an assumed shear strain distribution of parabolic shape
which has the same shear strain energy as the exact case from elasticity theory [13]. This theory

Page 32 af 125
applies to homogeneous beams and is called a single-variable theory because the transverse
deflection of the centerline can be expressed only in terms of the bending deflection.
Initially, the transverse deflection is assumed to comprise of a bending component and a shear
component as in equation (3-49). The assumed shear strain distribution is parabolic and given in
equation (3-96).
5 𝑦 2 𝜕𝑣𝑠 (3-96)
𝜀6 = ( − 5 ( ) )
4 ℎ 𝜕𝑥

Within the derivation, the static equilibrium equations for 2D elasticity, equation (3-78) and (3-79),
are applied and solved for to obtain the gross equilibrium equations of which equation (3-97) is
solved to yield the shear deflection in terms of the second derivative of the bending deflection as in
equation (3-98).
𝜕𝑀𝑥 (3-97)
− 𝑉𝑥 = 0
𝜕𝑥
𝜕 3 𝑣𝑏 5𝑏ℎ 𝜕𝑣𝑠
= 𝐸𝐼 − 𝐺
𝜕𝑥 3 6 𝜕𝑥

6 𝜕 2 𝑣𝑏 (3-98)
𝑣𝑠 = ( 𝐸𝐼 )
5𝐺𝐴 𝜕𝑥 2

Within the solution of equation (3-97) a term in equation (3-99) exist which has the third derivative
of the shear deflection in it but this equals zero for homogeneous beams, thus leading to the
dependency of only one variable.
ℎ (3-99)
𝑦=
2 1 𝑧 5 𝑧 3 𝜕 3 𝑤𝑠
∫ 𝑏ℎ𝑦𝐸 ( ( ) − ( ) ) ⋅ 𝑑𝑧 = 0
𝑦=−
ℎ 4 ℎ 3 ℎ 𝜕𝑥 3
2

When this theory is applied to a laminated beam in which the stiffness of each layer is weighted
differently by equation (3-100), the equivalent of equation (3-99) is given as equation (3-101)
which is not zero. This leads to it not being possible to express the shear deflection only in terms of
the bending deflection as shown in equation (3-102).
𝑛 𝑦=𝑦𝑘 (3-100)
(𝐴𝑖𝑗 , 𝐵𝑖𝑗 , 𝐷𝑖𝑗 , 𝐸𝑖𝑗 , 𝐹𝑖𝑗 , 𝐻𝑖𝑗 ) = ∑ ∫ 𝑄̅𝑖𝑗
𝑘
(1, 𝑦, 𝑦 2 , 𝑦 3 , 𝑦 4 , 𝑦 6 ) 𝑑𝑧
𝑘=1 𝑦=𝑦𝑘−1
𝑛 𝑧=𝑦𝑘 (3-101)
1 𝑦 5 𝑦 3 𝜕 3 𝑣𝑠
∑∫ 𝑏ℎ𝑦𝑄̅11
𝑘
( − ( ) ) 3 𝑑𝑧
𝑧=𝑦𝑘−1 4ℎ 3 ℎ 𝜕𝑥
𝑘=1

Page 33 af 125
𝑏 𝜕 3 𝑣𝑠 5𝑏 𝜕 3 𝑣𝑠
= 𝐷11 − 𝐹
4 𝜕𝑥 3 3ℎ2 11 𝜕𝑥 3

ℎ2 𝜕 2 𝑣𝑏 5 1 𝜕 2 𝑣𝑠 (3-102)
𝑣𝑠 = (−𝐷11 − ( 𝐹 − 𝐷 ) )
10𝐷55 𝑑𝑥 2 3ℎ2 11 4 11 𝜕𝑥 2

The main implication is that for a laminated beam, this theory does not lead to a single-variable
formulation but rather a higher-order theory which requires more degrees of freedom to be defined
at the boundary conditions to be solvable. Furthermore, the assumed displacement fields are not
correct for laminated beams and thus, a shear correction factor would be needed to display the
correct potential shear strain energy. For these reasons, no more work is done within this project to
develop this theory into a usable model for laminated beams, although the existence of a two
variable plate theory for orthotropic plates suggest that it may be possible to produce a usable beam
theory with two variables. [14]
By developing a finite element for a homogeneous beam by the direct stiffness method as in section
𝑑𝑢 𝑑𝑢
3.7, with the built-in boundary condition as 𝜙1 = (𝑑𝑦)𝑦=0 and 𝜙2 = (𝑑𝑦 )𝑦=0 and defining the
𝑥=0 𝑥=𝐿
18𝐸𝐼
shear constant as Φ = 𝐺𝐴𝑙2 , the elemental stiffness matrix takes the form in equation (3-103), which
𝑒
is similar to that which is derived for a Timoshenko beam without axial forces.
12 6𝑙𝑒 −12 6𝑙𝑒 (3-103)
2
𝐸𝐼 6𝑙𝑒 𝑙𝑒 (4 + Φ) −6𝑙𝑒 𝑙𝑒2 (2
− Φ)
3
𝐿𝑒 (1 + Φ) −12 −6𝑙𝑒 12 −6𝑙𝑒
2
[ 6𝑙𝑒 𝑙𝑒 (2 − Φ) −6𝑙𝑒 6𝑙𝑒 ]

In this theory, the expression for the shear force, when evaluated as the integral of shear stresses
over the cross-sectional area, equals that of the Timoshenko theory when the shear correction factor
4
is that of a homogeneous beam. This proves that the difference of a factor 3 in the effect of shear
deformation in the finite element theories is due to the definition of the built-in boundary
conditions, making this theory slightly more flexible. When compared to Timoshenko and exact
theory in [13], this theory more accurately predicts transverse deflection in the presence of built-in
boundary conditions but predicts the same deflection as the Timoshenko beam when the beam is
simply supported. The difference lies in the fact that both theories fix the vertical slope of the
middle of the cross-section as their built-in boundary condition but that a parabolic shear stress
distribution will have a higher rotation at the middle than an equivalent constant shear stress
distribution.
With the finite element models built with the proposed beam elements, the case of a simply
supported beam with a central loading is equivalent of each half of the beam being a cantilever
beam which is built-in at the middle and thus the case of 3 point bending does not yield the same
stiffnesses for these two theories. When comparing steel beams with a span of 400 mm, it can be
seen on Figure 13 that the calculated difference in stiffness is still relatively small but is rising with

Page 34 af 125
increasing importance of the shear deflection as the rotation of the cross-section at the beam
centerline increases.

Figure 13: Comparison of timoshenko and shimpi beam elements

This doesn’t in itself tell much about how severely a laminated Timoshenko beam with a flexible
core will be affected by the assumptions relating to the boundary conditions. Looking at the actual
shear strains of a laminated Timoshenko beam calculated from the 2D equilibrium equations, it is
seen that the shear strain is close to constant through the core which accounts for most of the
thickness of the beam. The shear correction factors calculated by equation (3-80) are also found to
be an order of magnitude smaller for the laminates in this project than the correction factor for a
homogeneous beam, thus indicating that the cross-section is approximately in a state of constant
shear strain. These arguments make it plausible that very little additional accuracy would come
from using a beam theory which more accurately represents the shear strain at the boundary
conditions and thus the Timoshenko beam theory will be used.

3.9 2D plane stress finite elements


While beam elements are formed from assumptions of the displacement, stress or strain field, 2D
elements can display an arbitrary two-dimensional field and thus include normal compression of the
core, include that the top and bottom composite skins have different curvatures and they can capture
the local fields near the supports and load applicators of the beam in 3 point bending.
Plane elements are elements which are either subjected to plane stresses or plane strains. Plane
stresses imply that the transverse thickness is negligible and plane strains imply that the transverse
thickness is infinite. As this thesis looks into assuming the plates in 3 point bending to be calculated

Page 35 af 125
as beams with negligible width, plane stress elements will be developed as isoparametric elements
as in [10]. A parametric element has the shape functions developed for a regularly shaped master
element and then the coordinates in the actual element, which may not have a regular shape, are
calculated as a multiplication of the shape functions by the nodal coordinates as shown in equation
(3-104). The internal displacements are mapped by shape functions as well as shown in equation
(3-105) and as the same shape functions are used for both, the element is said to be isoparametric.
𝑥 ∑𝒩 𝑥 (3-104)
{𝑦} = { 𝑖 𝑖 } = [𝒩]{𝑐}
∑𝒩𝑖 𝑦𝑖
𝑢 ∑𝒩 𝑢 (3-105)
{ } = { 𝑖 𝑖 } = [𝒩]{𝑑}
𝑣 ∑𝒩𝑖 𝑣𝑖

The shape functions are developed for the master element in the (𝜉, 𝜂)-coordinate system so as the
nodal displacements, {𝑑}, are given in the global coordinates, the strain-displacement matrix in
equation (3-106) includes transformations between these coordinate systems. To write these
equations more compact, derivatives are signified by subscript comma followed by the coordinate
which the derivation is with respect to.
Γ11 Γ12 0 0 (3-106)
1 0 0 0 Γ Γ22 0 0
[𝔅] = [0 0 0 1] [ 021 ]
0 Γ11 Γ12
0 1 1 0
0 0 Γ21 Γ22
𝒩1,𝜉 0 𝒩2,𝜉 0 𝒩3,𝜉 0 𝒩4,𝜉 0
𝒩1,𝜂 0 𝒩2,𝜂 0 𝒩3,𝜂 0 𝒩4,𝜂 0

0 𝒩1,𝜉 0 𝒩2,𝜉 0 𝒩3,𝜉 0 𝒩4,𝜉
[ 0 𝒩1,𝜂 0 𝒩2,𝜂 0 𝒩3,𝜂 0 𝒩4,𝜂 ]

Wherein the Jacobian matrix is given by equation (3-107), the Jacobian is given by (3-108) and the
inverse of the Jacobian matrix is given by (3-109), all replicated from [10]. The shape functions for
the plane quadrilateral elements Q4, Q8 and Q9 are given in Table 5 and the derivatives are given in
Table 6. The Q4 element is called a bilinear element with four nodes, the Q8 and Q9 elements are
quadratic elements with midside nodes and where the Q9 element also has an internal central node.
The Q8 element is called a Serendipity element and the Q9 element is called a Lagrange element.
𝑥,𝜉 𝑦,𝜉 ∑𝒩𝑖,𝜉 𝑥𝑖 ∑𝒩𝑖,𝜉 𝑦𝑖 (3-107)
[𝐽] = [𝑥 𝑦,𝜂 ] = [∑𝒩𝑖,𝜂 𝑥𝑖 ]
,𝜂 ∑𝒩𝑖,𝜂 𝑦𝑖

𝐽 = det([𝐽]) = 𝐽11 𝐽22 − 𝐽21 𝐽12 (3-108)

1 𝐽22 −𝐽12 (3-109)


[Γ] = [𝐽]−1 = [ ]
𝐽 −𝐽21 𝐽11

Page 36 af 125
Include only if node 𝑖 is present
𝑖=5 𝑖=6 𝑖=7 𝑖=8 𝑖=9
𝒩1 1 1 1 1
(1 − 𝜉)(1 − 𝜂) − 𝒩5 − 𝒩8 − 𝒩
4 2 2 4 9
𝒩2 1 1 1 1
(1 + 𝜉)(1 − 𝜂) − 𝒩5 − 𝒩6 − 𝒩
4 2 2 4 9
𝒩3 1 1 1 1
(1 + 𝜉)(1 + 𝜂) − 𝒩6 − 𝒩7 − 𝒩
4 2 2 4 9
𝒩4 1 1 1 1
(1 − 𝜉)(1 + 𝜂) − 𝒩7 − 𝒩8 − 𝒩
4 2 2 4 9
𝒩5 1 1
(1 − 𝜉 2 )(1 − 𝜂) − 𝒩
2 2 9
𝒩6 1 1
(1 + 𝜉)(1 − 𝜂2 ) − 𝒩
2 2 9
𝒩7 1 1
(1 − 𝜉 2 )(1 + 𝜂) − 𝒩
2 2 9
𝒩8 1 1
(1 − 𝜉)(1 − 𝜂2 ) − 𝒩
2 2 9
𝒩9 (1 − 𝜉 2 )(1 − 𝜂2 )
Table 5: Shape functions for quadrilateral elements. [10]

𝜕𝒩𝑖 𝜕𝒩𝑖 Include only if node 𝑖 is present*


𝜕𝜉 𝜕𝜂
𝑖=5 𝑖=6 𝑖=7 𝑖=8 𝑖=9
𝒩1 1 1 1 1 1
− (1 − 𝜂) − (1 − 𝜉) − 𝜕𝒩5 − 𝜕𝒩8 − 𝜕𝒩9
4 4 2 2 4
𝒩2 1 1 1 1 1
(1 − 𝜂) − (1 + 𝜉) − 𝜕𝒩5 − 𝜕𝒩6 − 𝜕𝒩9
4 4 2 2 4
𝒩3 1 1 1 1 1
(1 + 𝜂) (1 + 𝜉) − 𝜕𝒩6 − 𝜕𝒩7 − 𝒩
4 4 2 2 4 9
𝒩4 1 1 1 1 1
− (1 + 𝜂) (1 − 𝜉) − 𝜕𝒩7 − 𝜕𝒩8 − 𝜕𝒩9
4 4 2 2 4
𝒩5 −𝜉(1 − 𝜂) 1 1
− (1 − 𝜉 2 ) − 𝜕𝒩9
2 2
𝒩6 1 −(1 + 𝜉)𝜂 1
(1 − 𝜂2 ) − 𝜕𝒩9
2 2

Page 37 af 125
𝒩7 −𝜉(1 + 𝜂) 1 1
(1 − 𝜉 2 ) − 𝜕𝒩9
2 2
𝒩8 1 −(1 − 𝜉)𝜂 1
− (1 − 𝜂2 ) − 𝜕𝒩9
2 2
𝒩9 −2𝜉(1 − 𝜂2 ) −2(1 − 𝜉 2 )𝜂
Table 6: Derivatives of shape functions from Table 5

*The derivative of these additional terms are to be taken with respect to the same of either 𝜉 or 𝜂 as
the first term is.
With these and the width 𝑏, the stiffness matrix is calculated as the integral over the master element
given in equation (3-110).
1 1 (3-110)
[𝑘] = ∫ ∫ [𝔅]𝑇 [𝐸][𝔅] 𝑡 𝐽 𝑑𝜉 𝑑𝜂
−1 −1

Rather than evaluating this integral analytically, it is evaluated numerically using a quadrature rule.
The most commonly used rule are Gauss quadrature which samples internal points and multiplies a
weighing factor. Other quadrature rules also exist and also some exist which sample at the surface,
which is useful for beams and plates where the maximum bending stresses are at the surfaces.
The Gauss quadrature is evaluated at 𝑁 x 𝑀 points such that
𝑁 𝑀

[𝑘] = ∑ ∑ 𝜙(𝜉𝑖 , 𝜂𝑗 ) ⋅ 𝑊𝑖 ⋅ 𝑊𝑗
𝑖 𝑗

Where 𝜙(𝜉𝑗 , 𝜂𝑗 ) is [𝔅]𝑇 [𝐸][𝔅] evaluated at point (𝜉𝑖 , 𝜂𝑗 ) and 𝑊𝑖 and 𝑊𝑗 are weight factors. The
coordiantes of the evaluated points and the weighing factors are taken from Table 7.
𝑁 or 𝑀 Points 𝜉𝑖 or 𝜂𝑗 Weights 𝑊𝑖 or 𝑊𝑗

1 0 2
2 1
1
±√ ≈ ±0,5773502692
3

3 0 8
≈ 0,8888888889
9
3 5
±√ ≈ ±0,7745966692 ≈ 0,5555555555
5 9
4 18 + √30
3 2 6 ≈ 0,6521451548
±√ − √ ≈ ±0,3399810435 36
7 7 5

Page 38 af 125
3 2 6
±√ + √ ≈ ±0,8611363116 18 − √30
7 7 5 ≈ 0,3478548451
36

5 0 128
≈ 0,5688888889
255
10
±√5 − 2√ ≈ ±0,5384693101
7 322 + 13√70
900
≈ 0,4786286705
10
±√5 + 2√ ≈ ±0,9061798459
7
322 − 13√70
900
≈ 0,2369268850
6 ≈ ±0,2386191861 ≈ 0,4679139346
≈ ±0,6612093865 ≈ 0,3607615730
≈ ±0,9324695142 ≈ 0,1713244924
Table 7: Gauss Quadrature weight factors. [5]

For most purposes, 𝑀 = 𝑁 as the same degree of interpolation is used in both dimensions of the
plane element. A polynomial of degree 𝑝 is integrated exactly by having 𝑁 being the smallest
1
integer greater than 2 (𝑝 + 1) so to show a linear variation in strain (2nd degree polynomial of
displacements), a 2 by 2 quadrature is needed. The accuracy of integration rises by order of
integration but a too high order may overly stiffen higher-order displacement modes and thus the
accuracy of element behavior is not the same as accuracy of integration. For the Q8 and Q9
elements, full integration is 3 by 3 and reduced integration is 2 by 2.
The Q4 element is incapable of displaying pure bending due to the lack of quadratic terms in the
shape functions and will be overly stiff due to “shear locking”. To overcome this, the Q6 element is
formed by adding four generalized/nodeless DoF such that the displacement field is interpolated by
(3-111).
𝑢 ∑𝒩 𝑢 + (1 − 𝜉 2 )𝑎1 + (1 − 𝜂2 )𝑎2 (3-111)
{ }={ 𝑖 𝑖 }
𝑣 ∑𝒩𝑖 𝑣𝑖 + (1 − 𝜉 2 )𝑎3 + (1 − 𝜂2 )𝑎4

The added degrees of freedom in the Q6 element are responsible for incompatible modes which will
leave gaps or overlaps where elements meet and the element is made overly flexible but in such a
way that a refined mesh will converge towards the exact result. To generate the strain-displacement

Page 39 af 125
matrix of the Q6 element, that of the Q4 element is simply expanded with a matrix, [𝔅𝑎 ], which
operates on the nodeless DoF such that the strain-displacement matrix is given by (3-112).
[𝔅] = [[𝔅𝑑 ] [𝔅𝑎 ]] (3-112)

𝒩1,𝜉 0 𝒩2,𝜉 0 𝒩3,𝜉 0 𝒩4,𝜉 0 −2𝜉 0 0 0


𝒩1,𝜂 0 𝒩2,𝜂 0 𝒩3,𝜂 0 𝒩4,𝜂 0 0 −2𝜂 0 0
=
0 𝒩1,𝜉 0 𝒩2,𝜉 0 𝒩3,𝜉 0 𝒩4,𝜉 0 0 −2𝜉 0
[ 0 𝒩1,𝜂 0 𝒩2,𝜂 0 𝒩3,𝜂 0 𝒩4,𝜂 0 0 0 −2𝜂 ]

This Q6 element is however incapable of displaying constant stress if the initial shape is distorted
from a rectangular shape. To overcome this, the modified element named QM6 is formed by using
the values of [𝐽], 𝐽 and 𝑡 calculated at 𝜉 = 𝜂 = 0 for the calculation of [𝔅𝑎 ] but calculating [𝔅𝑑 ] as
usual. The stiffness matrix is then calculated as described in [15] by equation (3-113).
[𝑘𝑑 ] [𝑘𝑑𝑎 ]𝑇 (3-113)
[𝑘] = [ ]
[𝑘𝑑𝑎 ] [𝑘𝑎 ]
Where
𝑁 𝑀

[𝑘𝑑 ] = ∑ ∑[𝔅𝑑 ]𝑇 [𝐸][𝔅𝑑 ] ⋅ 𝑡 ⋅ 𝐽 ⋅ 𝑊𝑖 ⋅ 𝑊𝑗


𝑖 𝑗

𝑁 𝑀

[𝑘𝑑𝑎 ] = ∑ ∑[𝔅𝑎 ]𝑇 [𝐸][𝔅𝑑 ] ⋅ 𝑡𝑎 ⋅ 𝐽𝑎 ⋅ 𝑊𝑖 ⋅ 𝑊𝑗


𝑖 𝑗

𝑁 𝑀

[𝑘𝑎 ] = ∑ ∑[𝔅𝑎 ]𝑇 [𝐸][𝔅𝑎 ] ⋅ 𝑡𝑎 ⋅ 𝐽𝑎 ⋅ 𝑊𝑖 ⋅ 𝑊𝑗


𝑖 𝑗

If an element with internal nodes or nodeless DoF is used directly as-is, all DoF are transferred into
the global system to be solved for, even though some DoF are not acted upon by other elements. It
is common to condense the element into an equivalent stiffness matrix, acting only on the nodal
DoF on the boundary such that fewer DoF are transferred into the global system. The stiffness
matrix, DoF and reaction vectors are split into a retained part with subscript r and a condensed part
with subscript c as shown in (3-114).
[𝑘𝑟𝑟 ] [𝑘𝑟𝑐 ] {𝑑𝑟 } {𝑟 } (3-114)
[ ]{ }={ 𝑟 }
[𝑘𝑐𝑟 ] [𝑘𝑐𝑐 ] {𝑑𝑐 } {𝑟𝑐 }
By solving for {𝑑𝑐 } in the calculation of {𝑟𝑐 } and inserting into the calculation of {𝑟𝑟 }, an equation
of the equivalent stiffness matrix is found.
([𝑘𝑟𝑟 ] − [𝑘𝑟𝑐 ][𝑘𝑐𝑐 ]−1 [𝑘𝑐𝑟 ]){𝑑𝑟 } = {𝑟𝑟 } − [𝑘𝑟𝑐 ][𝑘𝑐𝑐 ]−1 {𝑟𝑐 } (3-115)

Thus, the equivalent stiffness matrix working only on the DoF at the boundary is given by (3-116).

[𝑘𝑒𝑞 ] = ([𝑘𝑟𝑟 ] − [𝑘𝑟𝑐 ][𝑘𝑐𝑐 ]−1 [𝑘𝑐𝑟 ]) (3-116)

Page 40 af 125
3.9.1 Validity of element formulations
The validity of the elements are tested by two methods. One is the patch test and the other is
modeling a beam with the elements and comparing against an analytical result.
To ensure that the elements function properly, they are subjected to a patch test as described in [10],
which is a numerical experiment in which the element has to be able to express a state of constant
stress, preferably also in a deformed shape. For elements where the patch test is not passed in a
deformed state, they are said to pass the weak patch test if a refinement of the mesh results in
convergence to a state of constant stress. Refinements are not tested for the patch test here.
To test the deformed element, a square is subdivided into four elements and the central node is
displaced. On Figure 14 below, the original shape is drawn in blue and the deformed shape is drawn
in red, exaggerated by a factor of 1000. The lower left node is fixed and the two other left nodes are
constrained horizontally while a force is applied at each node to the right, consistent with a constant
stress of 10 MPa.

Figure 14: Patch test of plane elements

The Q4 and QM6 elements passed this patch test by showing a uniform stress while the author of
this project was not able to program a functional Q8 or Q9 element.
To test the performance of these elements to show bending and shearing, these elements are
assembled to a steel beam with a length of 400 mm, a width of 275 mm and a thickness of 10 mm.

Page 41 af 125
The beam is supported at each end on the lower side while a central force is applied at the upper
side. This is shown on Figure 15 where the undeformed element is shown with blue stars at the
unloaded nodes and red stars at the loaded or supported nodes while the deformed shape are shown
with red circles at the nodes.

Figure 15: Plane element beam in 3 point bending

The number of elements lengthwise and through the thickness are adjusted such that the aspect ratio
are 1 and then more elements are added in both dimensions to add more through the thickness.
Having the elements be longer than they are thick would be more computationally efficient as less
elements would be needed but this distortion would also affect the accuracy of the computation so
to limit the scope of this project from also needing to investigate this, the aspect ratio of 1 are
chosen.
The stiffness of the beam is calculated and compared to the analytical result calculated by (3-117)
which is based on a Bernoulli-Euler beam and also against a Timoshenko beam calculated by
(3-118) which is given in [5].
48𝐸𝐼 𝑁 (3-117)
𝐾0 = = 3437,5 [ ]
𝑙3 𝑚𝑚
48𝐸𝐼 𝑁 (3-118)
𝐾1 = = 3430,8 [ ]
𝐸 ℎ 2 𝑚𝑚
𝑙 3 (1 + 𝜅𝐺 ( ) )
𝑙

As can be seen from the comparisons presented in Figure 16 and Figure 17, the QM6 element is
superior to the Q4 element in terms of displaying a stiffness that is closer to the analytical result but
both converge towards a higher stiffness than the theoretical values for beams. This was not
expected as the transverse compression and the supports being concentrated at the lower corners
was expected to yield a lower stress than the beams. This over-prediction of stiffness is assumed to

Page 42 af 125
be due to the elements being linear but this cannot be tested when the programming of functional
Q8 and Q9 elements have failed.

Figure 16: Comparison of plane elements to analytical beam Figure 17: Zoomed-in view of Comparison of plane elements to
solutions. Stiffness in Newton per millimeter of vertical analytical beam solutions
deflection of the loaded central node

As a compromise between accuracy and computation time, QM6 with 6 elements through the
thickness of the beam is chosen as it can be seen on Figure 18 that the QM6 element seems
reasonably converged while the Q4 element will require a much higher number of elements to
converge to a similar level.

Figure 18: Convergence of refinement of plane element beam

Page 43 af 125
3.10 Other 2D and 3D elements for laminate analysis
Besides the elements used here, there are other notable elements which can be used for analyzing
laminates.
Plate elements are two-dimensional elements which are formed by applying constraints to a three-
dimensional solid similarly to how beams are formed by applying constraints to a two-dimensional
plane stress region. These assumptions mean that the strain field through the solid can be calculated
from the deflections of the mid-surface, similarly to how the two-dimensional strains in a beam are
calculated from the deflections of the centerline. As the assumptions of a Bernoulli-Euler or
Timoshenko beam are based on the cross-sections being straight and are assumed perpendicular to
the centerline for the former but not the latter, the assumptions of plate elements relate to vectors
that are normal to the mid-surface. Kirchoff-Love plate theory, also called classical plate theory, are
analogous to Bernoulli-Euler beams in that these normals are assumed to remain perpendicular to
the midsurface after deformation, which imposes the constraint that the plate must be free of
through-thickness shear. Shear deformable plate theories are abundant like they are for shear
deformable beams but the most common theories assume constant shear strain through the thickness
and such the normals will remain straight but not necessarily perpendicular to the midsurface in the
deformed state. Often, this theory is called the Mindlin theory or the Mindlin-Reissner theory. [10]
The latter is a misunderstanding as the theory proposed by Reissner did not assume through-
thickness normal strain to be negligible as Mindlin did and thus the results differ between the
Mindlin and the Reissner plate theories even though they are very similar and were developed
around the same time. [16]
Plate elements are formulated as parametric elements similarly to how plane elements are formed
but there are three or five degrees of freedom per node depending on whether in-plane deformations
of the mid-surface are included. Apart from displacements, these are rotation of the normals about
the x and y axis, similar to how the rotation of the cross-section at the centerline is a degree of
freedom for beams. The similarity to beams is also found in that Kirchoff-Love plate elements are
𝐶 1 continuous as the slope of the mid-surface is equal between elements while Mindlin plate
elements are 𝐶 0 continuous as the slope of the mid-surface is not continuous across elements.
Mindlin plate elements are derived as the addition of a stiffness matrix for bending and a stiffness
matrix such that seperate integration schemes can be applied to the bending part than the shear part.
This is called selective integration and is used to combat a shear stiffening effect that these elements
have when the thickness is small. [10]
[𝑘] = [𝑘𝑏 ] + [𝑘𝑠 ] (3-119)

= ∫ [𝔅𝑏 ]𝑇 [𝐷𝑀 ][𝔅𝑏 ]𝑑𝐴 + ∫ [𝔅𝑠 ]𝑇 [𝐷𝑀 ][𝔅𝑠 ]𝑑𝐴


All of the elements will show too large flexibility when the plates are very thick but some element
formulations diverge to overpredicting the stiffness when the plates become very thin as shown in

Page 44 af 125
Figure 19. Only the Lagrange element under reduced or selective integration and the Heterosis
element match the theoretical stiffness for a large range of thicknesses. The Heterosis element is a
nine-node element where the transverse deflection degree of freedom are omitted at the centernode
and selective integration is used. Because the programming of functional plane Serendipity and
Lagrange elements failed in this project, plate elements are not developed either. As the shear strain
is assumed constant through the thickness of Mindlin plates and shells, a shear correction factor is
used which for laminated sections are calculated similarly to how [11] calculates it for beams but
just in two dimensions.

Figure removed due to Copyrights to be able to publish


Figure 19: Stiffness of plate elements compared to theoretical values. [10]

Shell elements are another category of elements which can be used for modeling laminates. Shells
have a lot in common with plates but may have initial curvature and even double-curvature which
the plates does not. This curvature couples bending and membrane stresses. Shell elements are
developed as isoparametric elements with plane stress but pose more challenges than plane and
plate elements as they can have both shear locking and membrane locking effects, which result in a
too-large stiffness. To counteract these locking phenomena, remedies such as reduced integration,
drilling DoF, incompatible modes and penalty functions are employed. [10][17]
Plate elements which include in-plane stresses are often referred to as flat shells. [10] This makes
shell elements more generally applicable and thus shell elements are used in commercial software
such as ANSYS whereas plate elements are not. [17]

Solid-Shell elements are solid elements which have been imposed some of the constraints
applicable to shells but have not been condensed into nodes on the midplane. Thus, the solid-shell
element have nodes with only translational degrees of freedom and is not assumed to have plane
stress and can display a three-dimensional state of stress. In ANSYS, this element is called
SOLSH190 and uses incompatible modes to increase accuracy and overcome thickness locking in
bending and it uses special kinematics to overcome locking when the element is very thin. [17]
While solid-shell elements can be formulated as a laminated section, ANSYS recommends against
using a single element for laminated structures as the stiffness can be highly overestimated and
instead, it is recommended to stack multiple layers of solid-shell elements. Solid-shell elements can
do much of the same as a regular solid element can but have the advantage of being able to model
both relatively thick and very thin plates and it is also much faster than solid elements. While a
proper comparison of computation times has not been conducted in this project, it is the experience
of the author that using SOLSH190 elements to model the core material of a sandwich plate is often
3-4 times faster than with SOLID185 with the same mesh size and approximately the same result.
Among the drawbacks for solid-shell element are limitations to the shape as the top and bottom
surfaces must be parallel and that the element is less developed in terms of nice features. The nice
features that ANSYS has for solid elements which are not applied to solid-shell elements include
automatic h-refinement to ensure convergence of results and non-linear adaptive remeshing.
An element type which is specially developed for analyzing laminates but does not currently exist
in ANSYS is the layerwise theory by Reddy [5]. In this theory, the in-plane properties are modeled

Page 45 af 125
by two-dimensional interpolations like the plane and plate elements while the through-thickness is
modeled through separate one-dimensional interpolations. This allows the in-plane mesh and the
through-thickness mesh to be refined separately and also, the calculations of the stiffness matrix of
a layer-wise element includes significantly fewer multiplications, additions and assignments than a
comparable 3D element. The layerwise theory does not apply an explicit assumption of the strain
through the thickness and instead, the through-thickness elements are only continuous in the
displacement (𝐶 0 -continuous) so that the strains may be discontinuous while the stresses are
continuous. This allows the theory to accurately display the “zigzag” stresses predicted by exact 3D
elasticity which also result in the bending stresses not varying linearly through the thickness of the
beam as it is assumed by equivalent single layer (ESL) beam and plate models. The model is said to
be partial layerwise when the discretization through the thickness is coarser than the amount of
lamina, in which case the layers that span multiple lamina are assumed homogeneous similarly to
equivalent single layer theories. The model is said to be full layerwise when the discretization
includes one or multiple layers through each lamina. Another feature of the full layerwise theory is
the ability to assume multiple displacement fields such that the boundary conditions can be forced
to be those of ESL theories to enable direct coupling to plate and shell elements or delaminations by
imposing a displacement that only applies to the layers above a delamination. While the
presentation of the layerwise theory in [5] is to solve a general 3D stress field, the layerwise theory
can also be applied to beams. [18]
As the implementation of layerwise theory as described in [5] requires successful implementation of
2D elements, it is omitted in this project but is noted as an area with a lot of potential for future
work.

3.11 Nonlinear analysis using ANSYS


In linear analysis, the stiffness matrix is formed with respect to the undeformed geometry,
deflections and strains are assumed small, the material properties are assumed constant and the
loads and boundary conditions are assumed constant. When any of these assumptions are not met,
the analysis is said to be nonlinear. To limit the scope of this project, nonlinearities are not
programmed into the elements that are calculated in Matlab and instead the commercial software
ANSYS is used to capture nonlinear effects.
Generally, there are three categories of nonlinearity, being geometric nonlinearities, material
nonlinearities and boundary condition nonlinearities.
3.11.1 Geometric nonlinearities
If a long and thin cantilever beam has a vertical load at the end it will deflect downwards under this
load. As the beam deflects downwards, cross-sections near the end will be rotated such that the
vertical load induces an axial strain over those cross-sections which is an effect that is nonlinear due
to the geometry. This axial strain will increase the bending stiffness of the beam in a way that is
more easily explained by an example with even less bending stiffness.
Assume a simply supported cable that is so slender that it has a negligible bending stiffness. This
slender cable now has a transverse load applied at its center. Assuming that the cable is initially
horizontal, it has practically no bending stiffness and if the deflections are calculated as a linear
analysis by dividing the load by the bending stiffness, the deflections will tend towards infinity.
Now if small strains but large deflections are assumed such that the deformed state of the cable

Page 46 af 125
forms the angle 𝜃 with the horizontal axis, the load at the center will form axial stresses along the
cable which resists further deflection. Thus it is shown that by including large deformations, the
calculation of infinite displacements can be avoided. In this case, the addition of a pretension of the
cable would add stiffness to the bending deflection, even when the cable is horizontal and the
addition of a compressive force would decrease the bending stiffness. The complete example is
given in lecture 8 of [19] to which the text book is [21].
To calculate these geometric nonlinearities in finite element analysis, there are generally two
methods that can be used.
The first method is the Total Lagrangian method in which a nonlinear strain measure, the Green-
Lagrange strain, is used along with a strain measure called the 2nd Piola-Kirchoff stress tensor to
describe the configuration at time t with respect to the original geometry such that the tangent
stiffness at time t can be calculated from the principal of virtual work which is then integrated over
the original volume. Within the derivations of the development of the Total Lagrangian method,
there are linearized equations which means that the solution at time 𝑡 + ∆𝑡 must be found in an
iterative manner within each load step to reduce the error that the linearization brings. The solution
at 𝑡 + ∆𝑡 are then used as a starting point for the iterative solution of the following load step, which
means that errors can accumulate if the solution did not converge before the next load step is taken.
The Total Lagrangian method involves transformations of stress concepts and terms which are
nonlinear in the displacement to make it possible to integrate over the original geometry and these
things make this method less used as it is simpler to refer to the geometry in time 𝑡 as is done in the
Updated Lagrangian method. ANSYS uses the Updated Lagrangian method so to keep the
relevancy, the Total Lagrangian method are not further described in detail.
The Updated Lagrangian method refers to the most recent calculated configuration which is at time
t. In the following, superscripts and subscripts to the left are employed to keep track of which times
the quantities are calculated and in with respect to which geometry such that 0𝑡𝜀𝑗𝑘 are the strain at
time t with the respect to the geometry at time 0. When no upper left subscript is written, it is
implied that this is the increment ∆𝑡. Also, the comma notation for derivatives is employed such
that a right subscript of comma followed by a letter means the derivative with respect to that letter.
The principal of virtual work at time 𝑡 + ∆𝑡 can be written in terms of the Cauchy stresses and
Cauchy strain, assuming small strains, and integrated over the unknown volume at time 𝑡 + ∆𝑡. It
can also be written in terms of the 2nd Piola-Kirchoff stress tensor and the Green-Lagrange strains
integrated over the volume of any known configuration but here it is chosen to integrate over the
known volume at time t.
(3-120)
𝑡+∆𝑡
∫ 𝑡+∆𝑡
𝜎𝑖𝑗 𝛿𝑡+∆𝑡 𝜀𝑖𝑗 𝑡+∆𝑡
𝑑𝒱 = ∫ 𝑡𝒮𝑖𝑗 𝛿 𝑡+∆𝑡𝑡𝑒𝑖𝑗 𝑡𝑑𝒱 = 𝑡+∆𝑡

𝑡+∆𝑡𝒱 𝑡𝒱

Where 𝒱 is the volume, 𝜎𝑖𝑗 is the Cauchy stresses, 𝜀𝑖𝑗 is the Cauchy strains, 𝒮𝑖𝑗 is the components
of the 2nd Piola-Kirchoff stress tensor, 𝑒𝑖𝑗 is the green-lagrange strains and 𝑡+∆𝑡ℛ is the virtual
work. Working with the principal of virtual work of the volume at time t and by defining the
nonlinear strain increment as (3-121) such that the Green-Lagrange strain increment can be
expressed by the addition of the Cauchy stress increment being linear and the nonlinear term
increment by (3-122). With these, the principal of virtual work becomes as in equation (3-123)

Page 47 af 125
where the right-hand side consists of the external virtual work and the internal virtual work
corresponding to the stresses at time t. [19]
3 (3-121)
1
𝑡𝜀̅𝑖𝑗 = ∑ 𝑡𝑢𝑘,𝑖 𝑡𝑢𝑘,𝑗
2
𝑘=1

𝑡𝑒𝑖𝑗 = 𝑡𝜀𝑖𝑗 + 𝑡𝜀̅𝑖𝑗 (3-122)

(3-123)
∫ 𝑡𝒮𝑖𝑗 𝛿 𝑡𝑒𝑖𝑗 𝑡𝑑𝒱 + ∫ 𝑡𝜎𝑖𝑗 𝛿 𝑡𝜀̅𝑖𝑗 𝑡𝑑𝒱 = 𝑡+∆𝑡
ℛ − ∫ 𝑡𝜎𝑖𝑗 𝛿 𝑡𝜀𝑖𝑗 𝑡𝑑𝒱
𝑡𝒱 𝑡𝑉 𝑡𝑉

This is so far valid for any amount of deformation but includes nonlinearities in the left-hand side
so for creating an approximation that is more easily calculated, the principal of virtual work is
linearized and is given in (3-124) where 𝐶𝑖𝑗𝑟𝑠 = 𝜆𝛿𝑖𝑗 𝛿𝑟𝑠 + 𝜇(𝛿𝑖𝑟 𝛿𝑗𝑠 + 𝛿𝑖𝑠 𝛿𝑗𝑟 ) is the fourth order
𝐸𝜈 𝐸
material tensor with 𝜆 = (1+𝜈)(1−2𝜈) and 𝜇 = 2(1+𝜈) for an isotropic material. For anisotropic,
laminated or nonlinear material laws, 𝐶𝑖𝑗𝑟𝑠 is entered appropriately.
(3-124)
∫ 𝑡𝐶𝑖𝑗𝑟𝑠 𝑡𝜀𝑟𝑠 𝛿 𝑡𝜀𝑖𝑗 𝑡𝑑𝒱 + ∫ 𝑡𝜎𝑖𝑗 𝛿 𝑡𝜀̅𝑖𝑗 𝑡𝑑𝒱
𝑡𝒱 𝑡𝒱

𝑡+∆𝑡
= ℛ − ∫ 𝑡𝜎𝑖𝑗 𝛿 𝑡𝜀𝑖𝑗 𝑡𝑑𝒱
𝑡𝒱

When (3-124) is discretized using interpolation functions by the finite element method, the left hand
side equals the virtual displacement increment of the nodes times the tangent stiffness matrix times
the actual displacement increment of the nodes. The right hand side equals the virtual displacement
increment of the nodes times the externally applied loads minus the force vector corresponding to
the internal stresses of the element. By dividing by the virtual displacement increment at both sides,
the result is equation (3-125) which must be solved iteratively to make the right hand side zero.
𝑡
𝑡𝐾 ∆𝑑 = 𝑡+∆𝑡
𝑅 − 𝑡𝑡𝐹 (3-125)

The right hand side can be thought of as residual of loads which in ANSYS Mechanical can be
viewed under the solution object as the force convergence in order to monitor how the solution
converges to a result within that load step. An example of the force convergence plot within
ANSYS is shown in Figure 20 where it can be seen that in the first iteration of each timestep (what
ANSYS calls substeps), the force imbalance is generally very high and it then iterates within a
timestep until the force imbalance is below a chosen criterion, after which the substep is converged
and the solution moves on to the next.

Page 48 af 125
Figure 20: Force convergence plot in ANSYS

Another convergence criterion which can also be seen in the solution object in ANSYS Mechanical
is the displacement convergence. The displacement is iterated in (3-126) such that at the 𝑘′𝑡ℎ
iteration of the 𝑖′𝑡ℎ loadstep, it is equal to the displacement at the previous iteration plus an
increment. The displacement convergence is then reached when the incremental displacement is
less than a value determined by the software. In equation (3-126), 𝑠 is a line search parameter that
ANSYS applies to stabilize the result by forcing smaller steps. 𝑠 = 1 by default and is reduced if
ANSYS notices convergence difficulties.
𝑡+∆𝑡 (𝑘) 𝑡+∆𝑡 (𝑘−1) (𝑘) (3-126)
𝑑𝑖 = 𝑑𝑖 + 𝑠∆𝑑𝑖
To solve (3-125), an iterative method such as the Newton-Raphson iterations are used. Many other
methods can be used, of which a modified Newton-Raphson method and an arc-length method are
available in ANSYS. The modified Newton-Raphson method is like the Newton-Raphson but
where the tangent stiffness matrix is only updated at the beginning of each load step. This requires
more iterations to solve but each iteration is faster as the tangent stiffness matrix does not have to be
recalculated. The arc-length method is a form of Newton-Raphson where instead of iterating by
going up a straight line to a certain load in each iteration, the straight line is followed up to an arc
some distance from the starting point of the load step such that the load-displacement curve can also
be followed downwards and into post-buckling regimes. [10][20]
By looking through the ANSYS documentation, it is not clear which iterative method is applied by
default in ANSYS Mechanical but it is likely that the program starts with the modified Newton-
Raphson and switches to the arc-length method if needed.
Within ANSYS Mechanical, enabling large deformations and large strains are simply done in the
analysis settings object by toggling the Large Deflection setting to “on”.
Large strains can be included in the updated lagrangian method by replacing the Cauchy
(engineering) strain by the Hencky (logarithmic) strain as these are identical for infinitesimal strains
[21 p. 615]. The engineering strain is not suitable for iterative calculations as adding multiple
incremental strains does not compute the same strain as going directly to the last strain level.
𝐿0 + ∆𝐿 𝐿𝑖 + ∆𝐿𝑖 (3-127)
𝜀𝐸 = ≠∑
𝐿0 𝐿𝑖

Page 49 af 125
More useful is the logarithmic strain, also sometimes called the true strain, which is defined as the
integral of incremental strains as shown in (3-128) for a one-dimensional strain and relates to the
engineering strain as in (3-129).
𝐿𝑖
𝑑𝐿 𝐿𝑖 (3-128)
𝜀𝐿 ∫ = ln(𝐿𝑖 ) − ln(𝐿0 ) = ln ( )
𝐿0 𝐿 𝐿0

𝜀𝑇 = ln(1 + 𝜀𝐸 ) (3-129)

When looking through the documentation of ANSYS, it is found that the elements are developed for
logarithmic strain but since this is equal to the engineering strain at infinitesimal strain levels, these
are automatically also useful for solutions where strains are assumed small.
3.11.2 Material nonlinearities
Many types of material nonlinearities exist but, in this project, elastoplasticity and laminate failure
are the most important and are the only material nonlinearities which will be treated.
Laminate failure is treated similarly as the fracture of an elastic body in the sense that if failure is
calculated at a load step, the stiffness corresponding to the failure mode of the failing lamina is
reduced or removed completely. After reducing the stiffness, the load step is recalculated to see if
any more failures occur and if not, the simulation continues with the reduced stiffness. ANSYS
supports fiber and matrix failures for laminated shells.
Elastoplasticity requires more sophisticated methods as it requires a yielding criteria, a flow rule
and a hardening rule. The yielding criteria is commonly set as the Von Mises criterion where a yield
strength is entered and this is also the criterion used for the bi- and multilinear hardening models
which can be chosen in ANSYS Workbench. The flow rule determines in which direction the
plastic strain occurs and with the bi- and multilinear hardening models in ANSYS, this is an
associative rule and the flow occurs normal to the yield surface.
Within ANSYS Workbench, there are several available choices for hardening models but only the
bi- and multilinear isotropic and kinematic hardening models will be treated here as these are the
simplest and are generally applicable to most homogeneous isotropic materials which the foam core
is assumed to be.
To have a bilinear model means that the stress-strain curve is assumed to consist of two straight
lines. The first line goes from zero and up to the yield stress with a slope that is the Young’s
modulus 𝐸 while the second line goes from the yield stress to infinity with a slope that is the
tangent modulus 𝐸𝑡 . The total strain is assumed to consist of an elastic strain 𝜀𝑒 and a plastic strain
𝜀𝑝 as shown on Figure 21.

Page 50 af 125
Figure 21: Schematic view of a bilinear model

The yield strength can be taken from data of the engineering stress since the strain at yield is often
small, meaning that the reduction of the cross-section due to the poissons ratio is also small. The
tangent modulus however must be estimated from a plot of the true stress versus the logarithmic
strain as the plastic strain will generally not be small. The true stress is the Cauchy stress evaluated
over the actual cross-section and not the initial cross-section as the engineering stress is. The actual
cross-section can be estimated during a tensile test by observing the deformations on the surface
using digital image correlation. [22] Often, such sophisticated measurement techniques are not
applied and the actual cross-section is calculated by assuming the elastic strains to be small and
assuming the plastic strain to occur under a constant volume, in which case, the true stress is
calculated from the engineering stress and the engineering strain by (3-130) [23][24].
𝐹 (3-130)
𝜎𝑡 = = 𝜎𝑒 (1 + 𝜀𝑒 )
𝐴
Since most materials have a gradual transition from the region of the stress-strain plot dominated by
elastic strains to the region dominated by plastic strain, the use of a bilinear model will have some
error. On Figure 22, some stress-strain curve is shown in black and possible bilinear models are
shown in green, red and blue. If the yield strength is chosen as the end of the linear region or the
stress at 0,2% plastic strain while the tangent modulus is taken as the slope within the region
dominated by plastic strain, the green graph is produced. With the green graph, the load-bearing
capacity is underestimated and the contained strain energy can be massively underestimated. If the
red graph is chosen to more accurately calculate the strain energy with a small overestimate during
the transition region, the yield stress may be highly overestimated. As the impact of each of these

Page 51 af 125
models depend heavily on how the actual stress-strain curve of a material looks like, the best fit of a
bilinear model is likely somewhere in between but a compromise could also be something like the
blue graph where the tangent stiffness is overestimated. As the complete stress-strain curve is rarely
available before purchasing a material and conducting tests on it, a graph like the blue graph could
be formed as a coarse approximation by looking at the yield strength, ultimate strength and strain-
to-failure which are often given in data sheets. By assuming that the maximum stress occurs at the
failure strain which is then calculated to be expressed as the logarithmic strain, the tangent modulus
can be found by connecting the yield stress and the maximum stress by a line. Depending on the
material, this would likely be an improvement over the green graph, but only for strains smaller
than the used failure strain. Limiting oneself to the failure strain given in data sheets may lead to a
very conservative estimate of the load-bearing capacity for ductile materials that exhibit necking
before failure in a tensile test. By using digital image correlation, to map the strains within the
necking band of steel specimen, it has been shown that the strain and stresses at failure can be much
higher than what is found by using an extensiometer only [24].

Figure 23: Multilinear hardening model


Figure 22: Possible bilinear models to fit a certain dataset

A multilinear model solves many of the issues that a bilinear model has but requires the stress and
strain to be known at multiple points after yielding and thus cannot be created from the data
commonly given in data sheets. The multilinear model, shown in Figure 23, employs multiple
sections with a constant tangent modulus and thus allows for the stress-strain curve to be replicated
at an arbitrarily fine resolution. The data that must be entered in ANSYS Workbench to create the
bilinear model are the stresses and the corresponding plastic logarithmic strain so the elastic strain
must be subtracted from the total strain to find the data that should be entered. After the last entered
datapoint in the multilinear model, ANSYS assumes a tangent stiffness of zero so it is important
that the last datapoint is at a higher plastic strain level than is expected to be seen in the simulation.
As the strain is often measured by an extensiometer as the average strain between two points, the
stress-strain curve at larger strain levels must be extrapolated from the tensile test data if there are
necking or if there are other reasons to believe that the material can be capable of exhibiting larger

Page 52 af 125
strain values than is observed in the tensile test. A suitable model could be Ludwik’s model (3-131)
which states that the stresses in the plastic region can be described as the yield stress plus a strength
coefficient 𝐾 multiplied by the plastic strain with the exponential of the strain hardening coefficient
𝑛. Fitting this equation to the plot of true stress versus plastic logarithmic strain before the
maximum stress allows for extrapolations into higher strain levels.
𝜎 = 𝜎𝑦 + 𝒦𝜀𝑝𝓃 (3-131)

Besides the choice of a bi- or multilinear model, a choice must also be made between isotropic or
kinematic hardening. The difference between these two models are how yielding alters the yielding
criterion for future yielding. For both hardening models, if the material is loaded until yielding at 𝜎𝑌
and then further loaded until 𝜎𝐵 before being unloaded, and then loaded back up, it will behave
elastically up to 𝜎𝐵 before any further yielding will occur. That is, the yield surface has changed but
the way that it changes is different for these two models. For isotropic hardening, the yield criterion
has expanded such that yielding in any direction now requires a yield stress of 𝜎𝐵 and the yield
criterion continues being isotropic under multiaxial loading. For kinematic hardening, the yield
criterion has moved to make the yield strength 𝜎𝐵 in the original direction but it has not expanded
so if the material is loaded in the opposite direction, the yield stress in that direction will be −(𝜎𝐵 −
2𝜎𝑌 ) so that the difference between the yield strength in two opposite directions are 2𝜎𝑌 . This
makes the yield behavior anisotropic in multiaxial loading after yielding. In reality, the material
behavior may be a combination of isotropic and kinematic hardening which can be included by use
of the Chaboche model but this is not further detailed in this project. [10] [20]

Figure 24: Idealized bilinear graph of isotropic and kinematic hardening.

In this project, the loading is not reversed and unloading of the specimen only occurs due to
laminate failures so the choice between isotropic and kinematic hardening is of lower importance.

Page 53 af 125
While the comparison between isotropic and kinematic hardening in Figure 24 is shown for a
bilinear model, it is similarly applicable to multilinear hardening models.
ANSYS supports bi- and multilinear isotropic and kinematic hardening for solids, solid-shells and
unlaminated shells but not for laminated shells.

3.11.3 Boundary condition nonlinearities


Boundary condition nonlinearities often refer to contact nonlinearities which occur when elements
come into contact with each other but also occur when elements are in contact through a non-
bonded contact. ANSYS deals with contacts by using contact elements to connect the surfaces in
contact. This element determines the distance between a contact pair either by the gauss points,
nodes or by a projected surface, where the surface projection method has the advantage of having
forces vary more smoothly and not jump if a gauss point or node leaves contact. The gauss-point
method is the default detection method in ANSYS. For transferring the forces, ANSYS has four
contact algorithms. [20]
The first is a pure penalty method, which means that the contact pair is allowed to penetrate each
other and the forces depend on how far they’ve penetrated and a normal and tangential stiffness.
The pure penalty method requires the least amount of iterations but are very sensitive to the contact
stiffness and may either allow large amounts of penetration if the stiffness is too low or lead to ill-
conditioned matrices when the stiffness is too high.
The second method is the pure Lagrange multipliers method in which no penetration is allowed and
chattering control parameters are used to stabilize the result but this requires more iterations and
adds degrees of freedom which increases computational cost.
The third method is the augmented Lagrange method which is the default contact formulation. In
this, the first two are combined so that the contact pressure is defined as the sum of the contact
stiffness times penetration and a Lagrange multiplier. This method is less sensitive to the contact
stiffness than the pure penalty method and requires fewer iterations than the pure Lagrange
multipliers method to determine the Lagrange multipliers.
The fourth method is also a combination of the two first but where Lagrange multipliers are used
for normal contact and the penalty method is applied to the frictional direction.
It is the experience of the author that contact nonlinearities lead to much more severe convergence
difficulties and increased computation times than geometric and material nonlinearities and such,
the use of contact nonlinearities should be avoided if possible. If the problem requires contacts to be
included, a number of settings within ANSYS Mechanical can be changed to improve convergence.
The contact stiffness is by default not updated within a load step but by changing this setting to
“each iteration” or “each iteration, aggressive”, the solution converges faster in some cases. If
convergence issues are still present, the normal stiffness may be reduced by changing normal
stiffness from “program controlled” to “manual” which brings up the option to adjust the normal
stiffness factor. The normal stiffness factor is 1 by default but by setting it to a lower value,
convergence may be reached faster but with more penetration.
Another issue is the discretization of curved surfaces and cylinders which may lead to initial

Page 54 af 125
penetration or initial gaps even when they are modeled to touch in the CAD software. This is solved
by changing the contact setting “interface treatment” to “adjust to touch”.
Another boundary condition nonlinearity is when a load is applied normal to the surface and the
structure is analyzed with large deflections. The direction of the load will then change with changes
to the orientation of the element.

3.12 Matlab FEM procedure – Boundary conditions


Prescribing degrees of freedom can be done in several ways. One simple method is the penalty
approach as described by [25] where the amendment of the 𝑖’th DoF by a prescribed value ∆𝑖 is
executed by multiplying a coefficient, 𝑐, to the stiffness of that DoF and adding 𝑐 ⋅ ∆𝑖 to the force
vector at that DoF as shown in (3-132).
𝐾11 𝐾1𝑖 𝐾1𝑛 𝑑1 𝐹1 (3-132)
[ 𝐾𝑖1 𝐾𝑖𝑖 + 𝑐 𝐾𝑖𝑛 ] { 𝑑𝑖 } = {𝐹𝑖 + 𝑐 ⋅ ∆𝑖 }
𝐾𝑛1 𝐾𝑛𝑖 𝐾𝑛𝑛 𝑑𝑛 𝐹𝑛

After the DoF has been calculated the reaction force is calculated by multiplying 𝑐 to the difference
in calculated and prescribed DoF value as shown in equation (3-139)
𝑅𝑖 = 𝑐(𝑑𝑖 − ∆𝑖 ) (3-133)

The magnitude of 𝑐 is commonly chosen to be 104 times the largest value found in the stiffness
matrix. The effect is similar to if the boundary was connected to ground by a very stiff spring and is
thus only an approximation as unwanted flexibility is introduced and choosing a very high value of
𝑐 may make the stiffness matrix ill-conditioned.
The penalty approach is not recommended for this problem for reasons which will become apparent
by running a parametric study of this method with 10, 100, 1000 and 5000 beam elements and with
penalty stiffnesses from 10 to 1019 . By varying the value of 𝑐 in the calculation of a steel beam with
a rectangular 275mm by 10 mm cross-section, lying flat, it is seen in Figure 25 and Figure 26 that
the choice of penalty stiffness is influential to the calculated stiffness. The calculated values for all
amount of elements are made in the same color as all the curves except those for the 500 element
model coincide until instability arises at 𝑐 = 1014 𝑁/𝑚𝑚.
The stiffness is calculated both directly from the prescribed displacement and reactionary load and
by correcting for the penalty by assuming the penalty stiffnesses to be springs in series with the
beam. By comparing the calculation using Bernoulli-Euler beam elements to the theoretical
48𝐸𝐼
stiffness , it is seen that the stiffness is only correctly calculated within 5 significant digits in the
𝐿3
range of 𝑐 = 1010 to 𝑐 = 1014 but that the correct stiffness can be calculated from the assumption
of linear springs in series at lower values of 𝑐 but only when the amount of elements are low as
instability is seen at both low and high values of 𝑐 with 1000 beam elements and that the theoretical
value is never calculated with 5000 elements. This calculation also shows that calculating 𝑐 as a
fixed factor times the highest value in [𝐾] is dangerous as refinement will then raise 𝑐 and possibly
introduce instability.

Page 55 af 125
Figure 25: Calculated stiffness versus penalty stiffness Figure 26: Calcualted stiffness versus penalty stiffness, zoomed
in

In these calculations, the displacement was solved by Matlab’s inbuilt matrix solver mldivide(K,F)
which may have affected the stability of the solution but nonetheless, the observations regarding the
penalty approach of boundary conditions makes it unsuitable for the problem in this report as it
introduces an error which is not negligible if 𝑐 is chosen incorrectly. Since a suitable value of 𝑐 is
dependent both on the global stiffness of the problem and of the stiffness of each element, ensuring
suitable values would require running parametric studies of 𝑐 for all laminates and would greatly
increase time spent and computational efforts.
An approach of prescribing the value of a DoF without introducing error as the penalty approach
does is the elimination method described by [25] where the row and column of the prescribed DoF
in the stiffness matrix are deleted along with the row of the prescribed DoF in the force vector. The
load components of all other DoF which are connected by the stiffness matrix to the prescribed DoF
are then subtracted the load corresponding to the prescribed DoF. In [10], without naming it the
elimination method, the same approach is described but in such a way that the size of the stiffness
matrix and force vector remains unchanged with the benefit that the force and DoF vectors are
complete with all of the values in the correct places. This formulation is as shown in (3-134) that if
the 𝑖’th DoF are prescribed 𝑑𝑖 = ∆𝑖 , the rows and columns corresponding to the 𝑖’th DoF are zeroed
out except for 𝐾𝑖𝑖 and in the force vector, other DoF are amended as in [25] and the 𝑖’th value are
set to 𝐹𝑖 = 𝐾𝑖𝑖 ∆𝑖 .
𝐾11 0 𝐾1𝑛 𝑑1 𝐹1 − 𝐾1𝑖 ∆𝑖 (3-134)
[ 0 𝐾𝑖𝑖 0 ] { 𝑑𝑖 } = { 𝐾𝑖𝑖 ∆𝑖 }
𝐾𝑛1 0 𝐾𝑛𝑛 𝑑𝑛 𝐹𝑛 − 𝐾𝑛𝑖 ∆𝑖
After the DoF has been calculated, the reaction forces are calculated from the original stiffness
matrix as in (3-141)
𝑛 (3-135)
𝑅𝑖 = ∑ 𝐾𝑖𝑘 𝑑𝑘
𝑘

Page 56 af 125
Yet another method which imposes boundary conditions without the error seen in the penalty
method is the use of lagrange multipliers where the stiffness matrix and force vector are kept
unedited and equations with additional unknown parameters, the lagrange multipliers, are added to
impose the constraints as shown in equation (3-136).
[𝐾] [𝐶]𝑇 {𝑑} {𝐹} (3-136)
[ ]{ } = { }
[𝐶] [0] {𝜆} {∆}
Where [𝐶], {𝜆} and {∆} are such that a prescription of the 𝑖’th DoF to ∆𝑖 by the 𝑗’th constraint
includes the added equations (3-143) and (3-144)
𝐶𝑗𝑖 ⋅ 𝑑𝑖 = ∆𝑖 (3-137)
𝑛 (3-138)
∑ 𝐾𝑖𝑘 𝑑𝑘 + 𝐶𝑖𝑗 ⋅ 𝜆𝑗 = 𝐹𝑖
𝑘
In the case of a simple prescription of DoF, 𝐶𝑗𝑖 = 1, but this formulation are also simple for
multipoint constraints or when a boundary condition is a roller on a slope. In this formulation, the
reaction force is easily extracted as 𝐶𝑖𝑗 ⋅ 𝜆𝑗 is the force imposed by the constraint.

To justify the choice of application of boundary condition these three presented methods are
compared to each other. The time it takes to apply boundary conditions, solve for the DoF and
calculate the applied load is recorded ten times for each method at various levels of refinement. To
limit the effect that of varying load from other programs on the computer, all three methods are
performed in the same program immediately after one another and timed by the “tic” and “toc”
commands in matlab. The timing may be different if it were to be run in another programming
language than matlab or if the implementation of programming the methods were done differently.
Also, the timing of each method may depend on the problem being analyzed so this is more of a
benchmarking of this particular implementation for use in this particular analysis than it is an
absolute benchmarking of the three methods.

Figure 27: Calculated runtime versus element count for Figure 28: Runtimes normalized against penalty method
different boundary condition methods runtime versus element count for different boundary condition
methods

Page 57 af 125
At an element count of 𝑛 = 10, there are such a large spread in the computation time in Figure 27
that it is difficult to draw any conclusions but at larger matrix sizes, it seems as shown in Figure 28
that the elimination method is faster than the lagrange method. For this reason and the fact that the
penalty approach was found to be inaccurate, the elimination method is chosen for prescribing DoF
at boundaries.

3.13 Matlab FEM procedure – Solver


Solving the equation [𝐾]{𝑑} = {𝐹} can be done in multiple ways, both directly and iteratively.
One way is to invert the stiffness matrix and multiply it to the force vector but this is a slow and
storage demanding solution as [𝐾]−1 is fully populated while [𝐾] is very sparsely populated. Other
methods for solving a system of linear equations directly are Gaussian Elimination, LU
decomposition, Cholesky’s method and Gauss-Jordan elimination as described in [26]. In Matlab, a
direct solver is implemented as an inbuilt function which has algorithm paths to choose a suitable
solver and also account for the sparsity of [𝐾] if it is input as a sparse matrix as described below.
For its ease of use and to limit the scope of this project, this solver which is called by mldivide(K,F)
or K\F is used as the default direct solver. Benchmarking mldivide(K,F) against multiplication of
[𝐾]−1 with beam elements shows that mldivide(K,F) is 1,8 to 6 times faster than inverting the
stiffness matrix when the stiffness matrix is smaller than 10000 by 10000 and sparse with the most
time saving being at larger sizes. This benchmarking is shown in Figure 29 and Figure 30.

Figure 30: Ratio of runtime between inversion and


Figure 29: Runtime versus element count for solving method
mldivide(K,F) versus element count

To further reduce the solving time, the stiffness matrix, which consists mostly of zeros can be saved
in the form where only the non-zero terms are saved. This is done by calling Ks=sparse(K) in
matlab where Ks is the resulting sparse stiffness matrix. By comparing the time it takes to create Ks
and solve the problem against solving with the full K matrix, it is seen that converting to a sparse
matrix is by far faster and that the benefit increases with problem size. This comparison, which the
result of is shown in Figure 31 and Figure 32 is made with elimination method boundary conditions
and with direct solving by mldivide().

Page 58 af 125
Figure 31: Runtimes versus element count for full matrix and Figure 32: Comparison of runtime versus element count of full
sparse matrix matrix and sparse matrix

3.14 Statistical analysis


In the data treatment in this project, statistical analysis is employed.
The section of data being examined is assumed to consist of an underlying function with a mean 𝜇
and a variance 𝜎 2 . With the section being analyzed starting at the a’th datapoint and ending at the
b’th datapoint, the mean 𝑦̅ and sample variance 𝑠 2 are calculated using the formulas found in
chapter 25.1 of [26].
𝑏 (3-139)
1 1
𝑦̅ = ∑ 𝑦𝑗 = (𝑦𝑎 + 𝑦𝑎+1 + ⋯ + 𝑦𝑏 )
𝑛 𝑛
𝑗=𝑎

𝑏 (3-140)
1 2 1
𝑠2 = ∑(𝑦𝑗 − 𝑦̅) = ((𝑦 − 𝑦̅)2 + (𝑦𝑎+1 − 𝑦̅)2 + ⋯ + (𝑦𝑏 − 𝑦̅)2 )
𝑛−1 𝑛−1 𝑎
𝑗=𝑎

The sample standard deviation is the square root of the sample variance 𝑠 = √𝑠 2 . To calculate the
interval of the mean with a conficence 𝛾 when the actual variance 𝜎 2 is unknown, the coefficient c
is determined from (3-141) by looking up the corresponding value in table A9 in appendix 5 of [26].
1 (3-141)
𝐹(𝑐) = (1 + 𝛾)
2
With the coefficient c determined, the coefficient k can now be calculated by (3-142) and then
confidence interval of the mean value is calculated by (3-143)
𝑐⋅𝑠 (3-142)
𝑘=
√𝑛
𝑐𝑜𝑛𝑓𝛾 {𝑦̅ − 𝑘 ≤ 𝜇 < 𝑦̅ + 𝑘} (3-143)

Page 59 af 125
Confidence intervals of variance can also be calculated. By looking up the values in the chi-square
distribution tables given as table A10 in [26] corresponding to equation (3-144) the values 𝑐1 and 𝑐2
are found with 𝑛 − 1 degrees of freedom.
1 1 (3-144)
𝐹(𝑐1 ) = (1 − 𝛾), 𝐹(𝑐2 ) = (1 + 𝛾)
2 2
Using the values of 𝑐1 and 𝑐2 , the limits of the confidence interval are calculated using equation
(3-145) and the confidence interval are then as in (3-146).
(𝑛 − 1)𝑠 2 (𝑛 − 1)𝑠 2 (3-145)
𝑘1 = , 𝑘2 =
𝑐1 𝑐2
𝑐𝑜𝑛𝑓𝛾 = {𝑘2 ≤ 𝜎 2 ≤ 𝑘1 } (3-146)

To assess whether there are differences between two specimen groups, hypothesis testing is
employed to test whether the mean of the two grous are are the same. The two dataset are assumed
to have unknown variances which are not necessarily equal and are tested with the matlab function
ttest2. In the ttest2 function, the value of t is calculated by equation (3-147) and compared to the
value in a t-table of the ‘Student’ distribution.
𝑥̅ − 𝑦̅ (3-147)
𝑡=
𝑠𝑥2 𝑠𝑦2

𝑛𝑥 + 𝑛𝑦
The effective degrees of freedom when two samples of unequal variance is being analyzed is
approximated by the Welch-Satterthwaite equation, (3-148), presented in [27].
2 (3-148)
𝑠 2 𝑠𝑦2
(𝑛𝑥 + 𝑛 )
𝑥 𝑦
𝜈′ ≈
𝑆𝑥4 𝑆𝑦4
+
𝑛𝑥2 𝜈𝑥 𝑛𝑦2 𝜈𝑦
Where 𝑠 is the standard deviation, 𝑛 is the sample size and 𝜈 = 𝑛 − 1 is the degrees of freedom.

Page 60 af 125
4 Method

4.1 Material choice


The composite laminate consists of a core with skin on either side made from laminated fiber which
is impregnated with resin. This resin may be spread onto dry fibers in the process of wet layup, it
may be sucked into the fibers with a vacuum pump during resin infusion or it may be impregnated
into the fibers before being shipped from the manufacturer, in which case it is considered pre-
impregnated, also called prepreg. Prepregs are often superior in performance to composites
manufactured with wet layup or resin infusion as it often has a more consistent fiber volume
fraction and a lower void content within each layer.
It is chosen to use prepregs in this project to limit complexity of manufacturing in hopes of more
consistent material properties and as the cost increase compared to purchasing dry fibers, resin and
hardener separately is deemed acceptable. The Gurit SE84LV prepreg system is chosen for being
curable at temperatures that the chosen foam core can withstand and without an autoclave. This
prepreg system exist with high elongation, intermediate modulus, high modulus or ultra-high
modulus carbon fibers but only the high elongation fibers can be purchased at the quantities needed
so this is what is chosen. A unidirectional prepreg is chosen as this allows for the optimal layup
being searched for with a high amount of fibers in the loaded direction.
The core can be any material but is most often a less dense material which is also more flexible.
Often balsawood, foams or honeycomb materials are used and while honeycomb materials of either
Nomex® or aluminum foil are much higher performance, these have complications in
manufacturing as the skins need to be manufactured separately and glued to the core whereas wood
and foams allow for the panel to made in one operation. Gurit corecell, balsaflex, Kerdyn Green
(PET) and PVC-HT can survive the temperatures of curing prepreg with 1 bar pressure. For their
homogeneous nature, foam cores are chosen for this project. Since the core is mostly loaded in
shear, the specific shear stiffness is sought to be as high as possible and thus the foams from Gurit
are compared in Figure 33.

Page 61 af 125
Figure 33: Comparison of foam core materials from Gurit. Specific shear stiffness has the units GPa/(kg/m3) while density has the
units kg/m3.

A trend is showing that the foam cores with higher density have higher specific shear stiffness with
the Gurit S-series being the stiffest, followed by the M-series which also span a higher range of
stiffnesses. Of the Gurit Corecell types, only the M-series are available at the regional reseller at the
time of purchase. In order to avoid having to purchase larger quantities than needed, the M80 foam
is chosen in 10 mm and 25 mm thicknesses while the M130 foam is bought in 30 mm thickness
which is then reduced in thickness to 25 mm by a planer.
To be able to test multiple parameters of bending and shear stiffness, five layups are chosen to be
𝑐 𝑐
tested by 3 point bending and perimeter shear stress test. These are [(0,90)2 , 21 ] , [(0,90)4 , 21 ] ,
𝑠 𝑠
𝑐1 𝑐2 𝑐3
[(0,90)7 , 2 ] , [(0,90)2 , 2 ] and [(0,90)2 , 2 ] where 0 represents layer of unidirectional SE84LV-
𝑠 𝑠 𝑠
HEC prepreg with the fibers oriented along the test direction of the 3PB test, 90 represents a layer
of prepreg at 90 degrees to the test direction of the 3PB test, (0,90)𝑘 represent 𝑘 repetitions of the
layup in the parenthesis, 𝑐1 represents a core of M80 foam with a 25mm thickness, 𝑐2 represents a
core of M80 foam with a 10 mm thickness, 𝑐3 represents a core of M130 foam with a 25mm
thickness and the subscript 𝑠 after the square brackets represent that these are all symmetric
lamiantes.

4.2 Specimen manufacturing


The prepreg sheets are cut to shape by first drawing the perimeter with a pen around an aluminum
template and then cutting with an electric scissor along the drawn lines. It is also possible to cut it
with a sharp utility knife but a cutting table of sufficient size wasn’t available. The width of the roll
of prepreg which has been bought is 400 mm so the 90-degree layers are cut from one strip that
long and another witch is 100 mm long so that 500 mm long 90-degree layers can be made. Cutting

Page 62 af 125
of the prepregs by this method is accurate within 2 mm but relies heavily on the skills and care of
the person cutting.
The core pieces is cut to shape with a table circular saw or a band saw as wood would be cut. The
M130 pieces is reduced from 30 mm to 25 mm by use of a planer meant for wood which leaves the
surface slightly rougher than as delivered. Cutting the core by this method has proven accurate
within one millimeter.
Before layup, a glass plate is cleaned and prepared either by taping a release film to it or by
applying a chemical release agent. At the manufacturing of the first few specimen in this project,
release film has been used which requires that the release film lies perfectly smooth against the
glass plate as any wrinkles or contaminants between the glass plate and the release film will transfer
to the surface finish of the part. When release agent is used, the plate must be cleaned with alcohol
or acetone and the plate must be scraped by the blade of a utility knife to remove any contaminants
or residue of tackytape, wax or resin that might not dissolve. After being thoroughly cleaned, the
glass plate is treated with a suitable release agent per the manufacturers guides. Here,
EasyComposites Easy-Lease release agent was used where a coat is spread with a cloth or paper,
given 10 seconds to start evaporating and then polished in with a second cloth until it has fully
evaporated. Between each coat, there is a waiting time of 15 minutes and after the final coat a 1
hour waiting time is had before the layup can begin.
For the first few specimen, the glass plate was prepared with a release film and the later test
specimen were made with a glass plate that was prepared with release agent. When only curing one
specimen, the release film method may be the most time-efficient although it still takes some time
to reduce the amount of wrinkles. When a larger amount of specimen are being cured, the release
agent is the most time-efficient as 5 or 6 glass plates can be prepared simultaneously so that the
waiting time in between each coating is utilized. Also, the release film would need to be reapplied
for each new layup while one treatment of release agent can be used for multiple cure cycles with a
single coat after demolding. The glass plate should also have tacky-tape applied along the edges for
the vacuum bag, which requires any release agent along the edge of the plate to be cleaned with a
suitable solvent as the tacky-tape will otherwise fail to stick to the glass when hot. The corners need
special care as these are often leaky. In Figure 35, a schematic view of one method of creating
corners are seen with the glass plate marked in green, the tacky tape along the edges marked in blue
and a supporting piece marked in red laid inside the corner. Often, corners are sealed simply by
putting one piece of tacky-tape over another but this can make a “tunnel” for the air to leak and thus
the method in Figure 35 seems superior for these glass plates.

Page 63 af 125
Figure 34: Glass plate with tacky-tape and one layer of prepreg Figure 35: Schematic view of application of
tacy-tape at a corner

The prepreg sheets are stacked on a prepared glass plate by removing the backing paper but leaving
the backing plastic on to keep the prepreg from distorting. A corner is then lined up with the layer
underneath and is lightly pressed stuck while lining up the edge which is parallel to the fiber
direction and lightly pressing that stuck as well. Holding the prepreg lightly tensioned, it is held
almost parallel to the layer underneath to ensure that the rest of the edges line up. If they do not, the
edge that was lightly pressed on can be made to release by quick yanks. If they do line up, the
prepreg is laid down, moving from one edge and across the fiber direction, rolling it on and pressing
out air bubbles by rolling a rigid roller, like the one seen on Figure 34, at a slight angle to the 90-
degree direction. When the whole sheet is attached, it is further pressed to the layer underneath by
rolling along the fiber direction with the roller while the backing plastic is still attached. To remove
the backing plastic, a piece of tape is attached to the backing plastic at a corner and given a fast
yank to get the backing plastic to release, at which point the plastic can be grabbed and pulled off.
As the 90-degree layer consist of one larger and one smaller sheet, the larger sheet should be placed
first to ensure that the fiber direction is as intended and the smaller sheet is laid after. To spread out
the effect of defects caused by small gaps or overlaps between the two 90 degree layers, the end at
which the small piece is laid is alternating. The core material is laid directly on top of the prepreg
and pressed on by hand as the resin within the prepreg will bond to the surface during curing. The
layup of the top sheets are done directly on top of the core in a similar fashion as for the bottom
layers. The bottom layers can be laid reasonably accurately when there are few layers but errors
accumulate with larger number of layers and it is also difficult to line up the top layers with the
bottom layers due to the core so a small angular error is to be expected. As the layers are aligned
across the length of 500 millimeters, the misalignment is assumed negligible since a 1-degree
misalignment would require one end to be off by 9mm compared to the other which was not
observed during layup.

Page 64 af 125
Figure 36: Laminate within a vacuum bag on a glass plate

After layup, a layer of release film is laid on top of the laminate, which may either be perforated or
unperforated. Using unperforated film will leave the surface slightly more shiny than when using
perforated as more resin is kept in the layup while perforated film will allow a small quantity of
resin to be sucked out while also removing any air that might otherwise be trapped between the
release film and the laminate. After the release film, wooden bars are placed along the edges of the
laminate to act as a mold, so the vacuum bag doesn’t pull the edges and corners down as much.
After this, breather cloth is laid over the laminate and down to the vacuum adapter so that air can be
sucked out through the breather before the vacuum bag is attached. While attaching the vacuum
bag, “ears” are added near the corners of the laminate, as can be seen on Figure 36, to make sure
that sufficient vacuum bag is added to form around the part. After closing the bag, a hole is cut for
the bottom part of the vacuum adapter which is left inside the bag, and the top part is screwed on
and vacuum is pulled. For these parts, the vacuum is kept below 100 mBar.
The composite specimens are cured in an oven where the temperature ramp is at most 2 degrees per
minute up to 100 degrees where it is held for 4 hours, after which the parts has cured.

The pieces of foam material that are meant for tensile testing are machined from a piece of foam
using a CNC mill to a shape according shape type 1A of DS/EN ISO 1798:2008. It is possible to
hotwire this material, but the surface partially melts, and the finish is not suitable for tensile testing
specimen.
The composite parts for material testing are cut from oversized cured laminates using a diamond
band-saw to have a section with approximately even thickness and are then wet sanded on the edges
with increasingly fine sandpaper while held in a holding tool to make sure that the sides are parallel
and without sanding marks. The top face of the specimens is also sanded to make the top and
bottom faces parallel and the compressive test specimen were also sanded on the ends while in a
holding tool. The tensile specimen in the longitudinal direction are manufactured according to ISO
527-5 so that they have a length of 250 mm, a width of 15mm and a thickness of 1mm. The tensile

Page 65 af 125
specimen in the transverse direction differ from the standard by being only 130 mm long and 18
mm wide due to the available prepreg strips that are left over from cutting for the 3PB and PSS test
specimen. Onto the tensile specimen, a steel tab with a chamfered end are bonded to both sides at
both ends using structural adhesive.
The compressive specimens are manufactured somewhat in accordance with ASTM D695 but as
rectangular pieces with a nominal width of 12,7 mm and a length of 80 mm.

4.3 3 point bending test


To prove equivalency, the formula student rules requires a three-point bending test to be made on a
specimen that is 500 mm long and 275 mm wide. The span between the supports are required to be
at least 400 mm and the specimen is required to be bent to a deflection of 12,7 mm by a central load
applicator with a radius of 50 mm. The deflection of 12,7 mm is assumed to be measured from
contact between the load applicator and the test specimen and 12,7 mm movement of the crosshead
from there, without accounting for the flexibility of the testing rig. The testing rig consist of a
Zwick Roell Z050 tensile/compression testing machine and a custom made combined 3-point
bending and perimeter shear stress test fixture. The test is conducted at a controlled rate of 5 mm
per minute.
The test fixture in the 3-point bending configuration is shown as a drawing on Figure 37 and seen in
pictures in Figure 38 and Figure 39. The test fixture has an adjustable span which is chosen to 400
mm so that a large deflection can be reached without risking that the specimen falls off the supports.
The supports are not regulated so here they are made from POM with a radius of 38 mm, supported
by a steel plate and raised by steel shims to have the sufficient clearance. 3D printed alignment tools
are used to ensure that the specimen is aligned with the testing direction.
To capture the load-displacement curve beyond initial failure for possible nonlinear modeling, the
specimen are loaded until a displacement of 50 mm as measured by the crosshead of the testing
machine.

Page 66 af 125
Figure 37: Drawing of the 3-point bending test fixture

Figure 39: 3-point bending test setup seen from the end
Figure 38: 3-point bending test setup seen from an angle

After conducting a bending test, the stiffness, yield strength, energy absorption and maximum load
within 12,7 mm must be extracted. The energy absorption is found by summing the force times

Page 67 af 125
distance increment of all the datapoints and the maximum load is simply the largest load in the data
where the displacement is less than the limit. Finding “the” stiffness as a single value of something
that may be nonlinear in nature and finding the yield strength of something that may not exhibit
yielding are challenging tasks.
As an interpretation of these requirements, it is assumed that an interval of data exists in which the
slope of the load-displacement diagram is approximately constant and that the yield strength is at
the end of this interval. Intervals beginning in all datapoints are examined and the mean slope and
confidence interval of the mean slope is calculated for this interval before extending the interval by
one datapoint and calculating the confidence interval of the mean slope of the new longer interval.
If the new interval has a smaller value of 𝑘 as calculated by (3-142), this interval is taken as the best
interval with that starting point and if not, the calculated values for this interval are discarded before
extending the interval further one datapoint. If 20 consecutive extensions of intervals have resulted
in increasing values of 𝑘, then it is assumed that the confidence interval will not improve by
extending the interval further and the start point of the interval is moved one datapoint and the
process is repeated. From this, the interval of approximately constant stiffness can be identified as
the interval with the lowest value of 𝑘 within the region that appears linearly elastic. This method of
determining the stiffness is very computationally expensive but is the chosen method for three-point
bending tests.

4.3.1 Test setup stiffness correction and baseline tube test


As the displacement is measured by the crosshead and the general design of the test fixture is not
regulated, the rules requires that the stiffness of the testing machine is corrected for by testing a
specimen for which the theoretical stiffness can be calculated. The measured stiffness is assumed to
be due to the test rig and the test specimen being two idealized springs in series as shown in
equation (4-1).
−1 −1 −1 (4-1)
𝐾𝑡𝑜𝑡𝑎𝑙 = (𝐾𝑠𝑝𝑒𝑐𝑖𝑚𝑒𝑛 + 𝐾𝑟𝑖𝑔 )

The rules require that the reference specimen are two baseline tubes as seen on Figure 40 which are
assumed to deflect only due to the bending and the theoretical stiffnesses of the tubes are assumed
to be as per (4-2).
48𝐸𝐼 ⋅ 2 (4-2)
𝐾𝑆𝐼𝑆 𝑡ℎ =
𝑙3
In the bending of baseline tubes shear deformation, cross-sectional deformations and local impacts
are assumed negligible but during bending, a local deformation of the tubes where the load is
applied quickly evolved as seen on Figure 41 and the cross-sections were also seen to deform into
an oval shape. As the deflection of the tubes is measured by the contact on the top of the tubes,
these deformations cannot be neglected as they introduce flexibilities which are counted towards the
test rig when the baseline tubes are used as reference specimen. Furthermore, the load of the tubes
onto the supports are much more concentrated than when plates are tested, which will naturally
result in the test rig being less stiff when testing tubes than when testing plates.

Page 68 af 125
Figure 40: Baseline tubes in 3-point bending Figure 41: Local deformation of baseline tubes

The load-displacement graph of these baseline tubes are shown in Figure 42 where it seems to have
a long approximately linear region but looking at the stiffness graph in Figure 43, it can be seen that
these tubes do not have any distinct region of constant stiffness as the stiffness is initially rising as
dominated by contact nonlinearity before dropping due to the deformations of the cross-section and
plastic nonlinearities.
𝑁
The measured stiffness of the baseline side impact structure tubes is 𝐾𝑡𝑒𝑠𝑡 𝑆𝐼𝑆 = 1911,9 [𝑚𝑚],
calculated as the mean of the interval from 0,5800 mm to 2,0558 mm and the yield load at 2,0558
mm is 3767,7 Newton. The absorbed energy after half an inch of displacement is 85,47 Joule and
the maximum loading is 8735,6 Newton. With this measured stiffness of the tube and the theoretical
stiffness calculated per (4-2), the stiffness of the testing rig is calculated to be 𝐾𝑟𝑖𝑔 𝑆𝐸𝑆 =
𝑁
7089,4 [𝑚𝑚].

Page 69 af 125
Figure 42: Baseline tubes 3PB load-displacement graph Figure 43: Baseline tubes 3PB stiffness graph

For the purpose of comparing models of laminates to experimental result, a more accurate estimate
of the testing rig stiffness is needed. To determine this, a steel plate the same size as the laminates is
used. Due to the availability of materials at the time of testing, this was done by laying a 175x500
plate alongside a 100x500 plate, both with a thickness of 10 millimeters. The measured stiffness is
𝑁
2988,8 [𝑚𝑚], calculated as a mean between 3,324 mm and 6,477 mm. Assuming that the anticlastic
curvature is not restricted due to the width of the plates and assuming that the through-thickness
compression is negligible, the theoretical stiffness can be calculated using the Timoshenko beam
N
theory which gives a theoretical stiffness of 𝐾𝑡𝑖𝑚 𝑡ℎ = 3430,8 [mm] corresponding to a rig stiffness
N
of 𝐾𝑟𝑖𝑔 𝑎𝑐𝑡𝑢𝑎𝑙 = 23199 [mm].

Figure 44: Load-displacement graph of 3PB of 10 mm steel Figure 45: Stiffness graph of 3PB of 10 mm steel plate
plate

When both the actual testing rig stiffness and the stiffness as calculated by the SES are known, a
value for the SES can be calculated from simulated data by using the actual rig stiffness to obtain an
equivalent “measured” value and using the SES rig stiffness to calculate the laminate stiffness

Page 70 af 125
which the SES would use to evaluate compliance as by (4-3). Since the actual stiffness of the testing
rig is much higher than that which is calculated from the baseline tubes as required by the rules, the
stiffness requirement for laminates are effectively reduced.
−1 −1 −1 −1 (4-3)
𝐾𝑙𝑎𝑚 𝑆𝐸𝑆 = (𝐾𝑙𝑎𝑚 𝐹𝐸𝑀 + 𝐾𝑟𝑖𝑔 𝑎𝑐𝑡𝑢𝑎𝑙 − 𝐾𝑟𝑖𝑔 𝑆𝐸𝑆 )

−1 −1
This does have the implication that there’s a singularity when 𝐾𝑙𝑎𝑚 𝐹𝐸𝑀 + 𝐾𝑟𝑖𝑔 𝑎𝑐𝑡𝑢𝑎𝑙 equals
−1
𝐾𝑟𝑖𝑔 𝑆𝐸𝑆 which for these stiffnesses exist when the laminate stiffness is a bit over 10.000 Newton
per millimeter as seen on Figure 46. This stiffness is far beyond what would be useful for a formula
student car as, in order to obtain it, it would be far too heavy to be competitive.

Figure 46: Singularity in laminate stiffness for SES

4.3.2 Simulation method – Matlab


The laminates are being simulated using finite element analysis under the assumption that they are
beams. The reason for choosing to use finite element analysis rather than analytical methods are that
finite element analysis is more suited for extension to nonlinear analysis and that finite element
analysis can be used to create a beam model where the core is modeled by 2D elements to include
compression of the core and local deformations near the supports.
Model 1: Beam
The model is linearly elastic and the beam model solves the timoshenko beam equations exactly.
Two elements are used where the only degrees of freedom per node are the rotation of the cross-

Page 71 af 125
section and the transverse deflection. At the free ends, the transverse displacement is prescribed
zero to make a simply supported boundary condition while at the shared node at the center, the
transverse displacement is prescribed a unit displacement. The boundary conditions are applied by
the elimination method. The stiffness of the laminate is calculated from the force required to apply
the prescribed unit displacement at the center node. The stresses in the laminate and in the core are
calculated and are extrapolated until failure due to the assumption of linear elasticity. If a failure is
calculated to happen before a displacement of 12,7 mm, the strength is taken to be the load when
the failure occurs, and the energy is calculated depending on whether a lamina or the core fails first.
If a lamina is calculated to fail before the core, it is assumed that all stiffness is lost and the only
energy that is absorbed is that of elastic deformation up until first ply failure. If the core is
calculated to reach a Von Mises equivalent stress equal to the level where 0,25% plastic strain is
observed in tensile tests, the beam is assumed to deform until 12,7 mm under constant load from
then on.
The failure of lamina are calculated per the max stress criterion as the criterion is linearly related to
the stresses, which makes it simpler to extrapolate until failure. Using the max stress criterion rather
than the more sophisticated criteria is regarded as having little impact to the accuracy of the
prediction of failure as all failures in the experimental part of this project is observed to happen
after both geometrical and material nonlinearities have become unneglectable. Thus, the stresses
and energy absorptions calculated by these linear models should not be trusted more than as a rough
approximation.
Model 2: Beam model as skin, 2D plane elements as core
In this second model, the skin is modeled by Timoshenko beam elements with both vertical and
horizontal displacement degrees of freedom at the nodes along with rotation of the cross-section.
The core is modeled by QM6 plane elements and the beam is modeled in full, including 50 mm
overhang over the supports in each end. At 50 mm and 450 mm from the left end of the beam, the
nodes on the lower surface are prescribed a zero vertical displacement and at 250 mm, the node on
the upper surface are prescribed a zero horizontal displacement and a unit vertical displacement.
The calculation of the shear coefficient for the beam elements that are used for the skins does not
account for the shear traction being non-zero on the side of the beam element that faces the core and
thus, the shear stiffness of the skin is incorrectly calculated too flexible. The shear stiffness of the
beam as a whole is however mostly determined by the core and as seen on Figure 47, the shear
stresses are approximately constant through the thickness of the core, which is the expected result
when it is laminated with relatively stiff skins. The shear stress in the core is seen to drop with
increasing amounts of layers in the skin which means that this model does calculate the shear stress
as incorrectly as the sandwich model but compared to model 1, the shear stresses in the core are
calculated about 10% higher here.

Page 72 af 125
Figure 47: Shear stress in core of beam with a 2D plane element core. The two green elements in the yellow part are a plotting error
caused by all of the nodes having the same calculated shear stress.

The failure of lamina and yielding of the core are treated similarly as with model 1. To evaluate
whether the core yields, the Von Mises equivalent stresses are calculated across all elements of the
core as seen in Figure 48. The linear-elastic analysis of 2D elements with a point-load includes a
stress singularity which means that the stresses calculated near the loads will never be mesh-
independent although the beam elements for the skin will help distribute the load a bit. As a
consequence of this, the elements in the core near the nodes with prescribed DoF are excluded from
the calculation of stresses and it is assumed that 5 mm or 1 element (whichever is further) away, the
singularity is sufficiently dispersed.

Page 73 af 125
Figure 48: Von Mises equivalent stresses in the core of a beam with 2D plane elements for core

To find the optimum layup for a part of the car, for-loops are used to simulate all possible layup
combinations with the M-series core foams, at chosen increments of core thickness and up to 6-
layer thick skins. Simulating all possible combinations of laminate angles is not considered feasible,
even if the angles are chosen in increments as for 𝑛 layers with 𝑣 possible increments of angles,
there’s 𝑣 𝑛 possible combinations, making for millions of possible layups. Instead, an assumption is
made of the lamina angles in the optimum layup. It is assumed that the highest legal amount of
fibers is in the loaded direction, so the outermost layers are placed at 0 degrees, while the remaining
are placed at alternating +/- 10 degrees.
After calculation of the stiffness, strength and energy absorption, the laminates with sufficient
results are ranked according to their mass so that the laminate with the lowest weight can be chosen
for further analysis.
To do the optimization, the stiffnesses of the core material from the data sheets are used and the
limit for yielding is taken as the tensile strength from the data sheet as the observed 0,25% proof
strengths were above these values for the tested materials.
4.3.3 Simulation method – ANSYS
The plates are simulated by modeling a quarter plate and thus taking advantage of the symmetry
planes by use of symmetry regions in ANSYS. Two models are used, analogous to the beam models
described above.

Page 74 af 125
Figure 49: Boundary conditions of a shell element model (full plate shown)

The first model is using shell elements with a defined layered section as ANSYS calculates the
laminate properties as an equivalent single layer. The plate is modeled without any overhang past
the supports, with a displacement boundary condition at the supports and a remote displacement
boundary condition as the applied load.
In this model, geometrical nonlinearities are included and composite stiffness reduction is included
based on the Hashin failure criteria. The stiffness reduction of a failed lamina is set to 75% as a full
reduction of stiffness has shown to lead to unstable results with very early failure. Yielding of the
core is not included in this model as enabling the setting for layered sections causes ANSYS to
crash.

Figure 50: Full model boundary conditions. Quarter plate shown. Symmetry boundary conditions are on the faces and edges to
simulate a full plate.

The second model in ANSYS is the most sophisticated in this project and is referred to as the full
model. In this, the core is modeled by 8 solid-shell elements through the thickness while the skins
are modeled as equivalent single layers by shell elements. The elements are approximately square
with a width of 4,8mm. A displacement boundary condition is applied to a line on the bottom
surface and a flexible remote displacement is applied to the top surface to simulate the physical

Page 75 af 125
boundary conditions while symmetry regions are applied to the surfaces and edges to enable a
quarter plate to be modeled.
Overhang beyond the supports is included, geometrical nonlinearities are included, laminate
damage of the skins is included by a stiffness reduction of 75% based on the Hashin failure criteria
and yielding of the core is included in the model. The yielding of the core is modeled as bilinear as
attempts to apply a multilinear material model has not been successful within the time period of the
project.
The simulated data of most of the simulations with the full model show a nonlinear stiffening effect
before yielding that is not seen in the experimental results and thus the stiffness is calculated as the
average over the first millimeter of displacement.

4.4 Perimeter shear stress test


The perimeter shear stress test is required by the rules to show certain strength levels for certain
parts of the monocoque and to use for estimating required bracket size for bolted connections.
The test is conducted by pushing a cylindric punch with a diameter of 25 mm through a rectangular
sample with a width of 100 mm which is supported by a steel plate with a 32 mm hole underneath
the punch. The hole is visible in Figure 40 as it is integrated into the 3PB test fixture and the punch
can be seen in Figure 51.

Figure 51: Perimeter shear stress test

The first peak, which is to be used for extrapolating required bracket size is defined as the largest
force before the top skin is fully penetrated, regardless of any damage prior to this peak as shown
on Figure 52. For the sake of the SES, the failure is assumed to be due to pure shear along the
perimeter of the punch, so if the upper skin has a thickness of 𝑡1 which failed at a load 𝑃1 , then the
shear strength which is considered the same in all directions is calculated per equation (4-4).

Page 76 af 125
𝑃1 (4-4)
𝑆𝑠 𝑆𝐸𝑆 =
𝑡1 ⋅ 25 ⋅ 𝜋

Figure 52: Perimeter shear stress test peak detection

4.4.1 Simulation method – ANSYS


It has not been possible for the author to conceive of any simplifications to 1D or 2D that would
enable simulation of the PSS test results, so the proposed method is a nonlinear analysis using
ANSYS with a solid or solid-shell core. The skins can be modeled as either shell elements or solid-
shell elements and the core can be modeled either as solid or solid-shell elements.
Using shell elements for the skins have the benefit of being more computationally efficient and that
changing the layup does not require CAD modelling and thus are also faster. Using solid-shell
elements for the skins have the benefit of being able to show 3-dimensional states of stress and that
contact debonding can be added to the contact between the elements to enable delamination.
Using solid-shell elements for the core has the benefit that it runs faster than an equivalent mesh of
solid elements and that it has features to limit shear stiffening. Using solid elements for the core has
the benefit that the core can be remeshed by the application of a nonlinear adptive region if the
mesh becomes too distorted.
The symmetry of the test is exploited and only a quarter of the test specimen is modeled and
symmetry conditions are applied to the sides and edges of the skins and core. The underside of the
specimen is supported by a displacement boundary condition except for a circular region where the
hole in the test fixture is. The load is applied to a circular region by a remote displacement with a
coupled behavior as shown on Figure 53.

Page 77 af 125
Figure 53: ANSYS model of PSS test with solid-shell top skins

4.5 Tensile tests


The tensile tests are conducted with a Zwick Roell Z050 tensile/compression testing machine with a
tensile testing fixture that utilizes wedge grips to hold the specimen. The specimens are centered so
that the centerlines of the specimens are coaxial with the testing machine grips and are pulled at a
controlled rate until failure. The composite test specimens are pulled at a rate of 0,5 mm/min while
the foam specimens are pulled at a rate of 1 mm/min. An external extensometer is used to measure
the elongation between two points and the strain is assumed constant within the measured length.
The gage length is adjusted to be within one millimeter of the target, as measured by a caliper and
by drawing lines on the specimen. The foam specimen is tested with a gage length of 40 mm while
the composite specimens are tested with a gage length of 50 mm.
The specimen cross-sectional width and thickness are measured with a micrometer at 5 locations so
that an average initial cross-section can be calculated for stress calculations.

4.6 Compressive tests


The compressive tests are conducted on a Zwick Roell Z050 compression/tension testing machine
with flat compression tools on and with a thin specimen holding tool per ASTM D695 as seen on
Figure 54.
To measure strain accurately on compressive test specimens, it is recommended to use strain gages
on untabbed specimens and to measure the strength accurately, it is recommended to use tabbed
specimens where the ends of the tabs and specimen are machined to be flat at the end. Due to the
workload in preparing the specimens to these recommendations, the specimens are manufactured as
plain rectangular specimen and are tested without strain gages so that approximate results may be
found.

Page 78 af 125
The compression of the specimens are measured by the translation of the crosshead of the test
machine and to account for the flexibility of the test machine, the tested specimen and the test
machine are assumed to be two ideal springs in series. To find the stiffness of the test machine, steel
specimen are tested and the stiffness of the machine is calculated similarly as done for the 3 point
bending tests by assuming the tested specimen and the machine to each be an ideal spring in series.
Four steel specimens are tested of which two are discarded due to the ends being not sufficiently
flat and the stiffnesses of the other two are found by searching for a region of approximately
constant stiffness, similarly as done for 3-point bending specimen.
Tested stiffness Theoretical stiffness Test machine stiffness
N/mm N/mm N/mm
Specimen 3 20.009 +/- 115 65.774 28.757 +/- 239
Specimen 4 19.420 +/- 125 65.574 27.592 +/- 252

The stiffness of the machine is not uniquely determined by these two specimens but is assumed to
be the mean of the calculated stiffness and thus a machine stiffness of 28.174 N/mm is used to
correct the calculated stiffnesses of the composite specimen. This is done by calculating the
displacement of the idealized spring that is the machine and subtracting it from each data point.

Figure 54: ASTM D695 compressive test fixture

4.7 Fiber volume fraction analysis


To determinine the fiber volume fraction, test specimens are manufactured in strips which are then
cut into smaller rectangles that are used in a burn-off test similar to ASTM D2854. All the test
specimens are cured at 100 degrees for 4 hours under a vacuum of 90 mBar.

Page 79 af 125
Three groups of test specimens are manufactured. One group is 3-layer specimens with a nominal
thickness of 1mm cut from excess length of the tensile test specimens, cured with perforated release
film and breather which absorbs some excess resin to improve the fiber-volume ratio. The two other
groups are made from a single layer of prepreg with a nominal thickness of 0,3mm. One group is
cured with perforated release film like the thicker specimens and the other group was cured with
unperforated release film to evaluate the influence of absorbing excess resin.
After cutting, the edges of the specimen are sanded clean and then the specimens are weighed to
find the mass 𝓂𝐿 . The sides are then sealed with clear nail polish and left to dry before the test
specimen are then weighed again to get 𝓂𝑑𝑟𝑦 and then submerged into water to be weighed yet
again to get 𝓂𝑤𝑒𝑡 .
The specimen are then burned in an oven at temperatures between 500 and 600 degrees Celsius to
evaporate the resin, leaving only the fiber behind.
Burn-off tests are often not recommended for carbon or polymer fiber as the fiber can also burn off
but the timing and temperatures used here were taken from a test procedure created by the company
Space Composite Structures Denmark (SCSDK), with whom the university shares the composite
lab, as these have tuned this procedure to yield accurate results. As the test procedure is the
intellectual property of SCSDK, the timings and temperature used are not disclosed but worth
noting is that due to the small weight of the tested specimens, all of them are in the same weight
bracket and burned for the same duration despite one group of specimens weighing about 3 times
that of the others. When burn-off tests are used despite the risk of fiber burn off, it is because it is a
simple, cheap and relatively fast method of measuring the fiber volume fraction. Other methods
include x-ray computed tomography and the matrix digestion method where the matrix is
chemically eroded or dissolved.
From the dry and wet weights including nail polish and the density of water, the volumes of the
specimens are calculated as the difference in weight divided by the density of water as given in
equation (4-5).
𝒱𝑑𝑟𝑦 − 𝒱𝑤𝑒𝑡 (4-5)
𝑉0 =
𝜌𝑤𝑎𝑡𝑒𝑟
The density of the laminate is then calculated by (4-6) from the specimen volume and the weight
excluding nail polish, from the assumption that the volume of nail polish is negligible.
𝓂𝐿 (4-6)
𝜌𝐿 =
𝒱0
Taking the density of the fiber, 𝜌𝑓 , and matrix 𝜌𝑚 from manufacturers data sheets, the fiber volume
fraction is calculated by (4-7) from the weight of the remaining fiber after burning.
𝓂𝑓 𝜌𝐿 (4-7)
𝓋𝑓 = ⋅
𝓂𝐿 𝜌𝑓
The fiber and matrix weight fractions are then calculated from (4-8) and (4-9) respectively.
𝓂𝑓 (4-8)
𝓌𝑓 =
𝓂𝐿

Page 80 af 125
𝓂𝐿 − 𝓂𝑓 (4-9)
𝓌𝑚 =
𝓂𝐿

The void (volume) ratio is calculated by (4-10) using the calculated density and the theoretical
density as calculated by (4-11) with the calculated weight fractions and the material densities from
the data sheets.

𝓌𝑟 𝓌𝑓 (𝜌𝐿,𝑡ℎ𝑒𝑜𝑟𝑒𝑡𝑖𝑐𝑎𝑙 − 𝜌𝐿 ) (4-10)
𝓋𝑣 = 1 − 𝜌𝐿 ⋅ ( + )=
𝜌𝑟 𝜌𝑓 𝜌𝐿,𝑡ℎ𝑒𝑜𝑟𝑒𝑡𝑖𝑐𝑎𝑙
1 (4-11)
𝜌𝐿,𝑡ℎ𝑒𝑜𝑟𝑒𝑡𝑖𝑐𝑎𝑙 = 𝓌 𝓌𝑚
𝑓
𝜌𝑓 + 𝜌𝑚
When the fiber and void volume fractions have been calculated, the matlab function ttest2 is used to
perform hypothesis testing on whether the means of two dataset are equal.

Page 81 af 125
5 Results
In this section, the main findings of the report are presented. The first result to be presented is the
thickness of the cured laminas. After testing, specimen 1 is cut with a diamond bladed saw to reveal
cross-sections of the laminate which are then looked at under a microscope. Using a digital camera
on the microscope and the accompanying software, digital rulers are placed on the picture and it is
seen that the laminates are consistently very close to 0,3mm in thickness. This also matches that the
total thickness of a 4 layer skin is measured with a caliper to be approximately 1,2-1,25mm.

Figure 55: Cut surface showing the cross-section of a skin consisting of 4 layers of CFRP in alternating 0 and 90-degree directions

5.1 Defects
Defects such as voids and contaminants exist within laminated parts and although care can be taken
to minimize these defects, they cannot be avoided completely.
5.1.1 Contaminants
The layup has not been done in a clean room but in a composite working lab, which has introduced
some contaminants. To assess the amount of contaminants, the first 3-point bending test specimen
has been examined on the side facing the glass plate by shining a bright light onto the surface and
circling and counting all visible contaminants as on Figure 56. This does not tell about the amount
of contaminants within the laminate but it may give insight to the importance of cleanliness.

Page 82 af 125
By circling and counting, 148 pieces of solid contamination has been identified, ranging from
millimeter long fibers to centimeters long fibers. The most abundant type of contaminant is very
short fibers of very low diameter which is curled up so much that they are assumed to be very
flexible and thus probably are fibers from either clothing or breather material. Other contaminants
appear to be human hairs and also what appears to be glass fibers can be seen. As the layup of this
specimen has been done on release film rather than on a glass plate with release agent, it is possible
that some surface imperfections due to wrinkles in the release film have been misidentified as long
fibers. This possible misidentification would be a small amount of the counted defects since the
imperfections due to wrinkling are only seen along the fiber direction and often looks different from
the contaminants as seen on Figure 57.

Figure 56: Contaminants catching the light when seen a bit Figure 57: Contaminants in circles. Imprints of wrinkles of the
from the side release film seen as horizontal lies that catches the light

5.1.2 Voids
To examine for voids, the sides of the tensile test specimens which has been wet sanded with fine
grade paper are being looked at under a microscope. Here, two sizes of voids are seen. Small voids
that are common and are likely to not be possible to prevent and rarer large voids which may be
possible to prevent by exercising extreme care to avoid air traps during layups. The small voids
were seen to have a preference for one direction, such that parallel to the fibers, the length of the
voids are measured to be up to 2mm with 1mm being the most common size. When inspecting by
eye, they do seem to be longer, but when looking under the microscope, it seems likely that this is
due to large number of voids in the interface between layers and also due to the sanding being
parallel to the fibers. Two voids, marked with rulers, are visible on Figure 58 although the
resolution and lighting of the image makes the cracks stand out less than when they are seen
through the eye-piece. When looking onto a plane perpendicular to the fibers, the voids seem
circular with a diameter of 10-20 micrometers and are seen to exist in large numbers in the
boundary between layers as seen on Figure 59 and Figure 60.

Page 83 af 125
Figure 58: Voids seen when looking parallel to the fibers. The rulers are 2mm and 0,5mm long.

Figure 59: Voids photographed with microscope built-in camera, looking onto the ends of the fibers. Voids marked from the left with
three 20 micrometers voids and two 10 micrometer voids.

Figure 60: Voids looking onto the ends of the fibers, photographed with an external camera through the eye-piece. Note that the
uneven top surface is filled with fibers and are not just pooled resin.

Page 84 af 125
The larger voids, where a large portion of air is trapped are rare with observations on only two of
the 25 tensile test specimen and one observation on 15 compressive test specimens. The large voids
seem to be about 4mm long, 0,2mm in the thickness direction and with unknown depth in the
width-direction of the specimen. While it does seem that the thickness-dimension is much smaller
than the other dimensions, it is not clear whether there are any preferences for the extension in the
two in-plane directions. On one of the specimens which had large voids, a larger than usual number
of smaller voids was also observed, which may indicate that this specimen had not been rolled with
sufficient pressure during the lay-up. This is shown on Figure 62.

Figure 61: Large void seen parallel to the fibers direction.

Figure 62: Large void seen when looking onto the 90-degree direction.

5.1.3 Large-scale surface unevenness


On the 90-degree tensile test specimen, valleys and peaks in the thickness are observed with a high-
thickness region just next to a low-thickness region with about 0,2 mm difference in the specimen
thickness. These valleys and peaks span the whole width of the specimen and are parallel to the
fiber. The apparent radius seems to match that of the roller that was used to roll the layers onto each
other, which would suggest that the pressure has been too high at times where the rolling direction
was parallel to the fibers. When looking under a microscope, it is seen that the fibers extend up into
most of the surface unevenness.
This suggests that great care should be taken to avoid pressing too hard with a roller if the roller is
parallel to the fibers which poses a dilemma as this is the way that it is intuitive to roll the layers
onto each other during lay-up to attempt to avoid trapping air. One of the specimens that had a lot of

Page 85 af 125
micro voids and a region with large voids was among those that had large-scale surface unevenness,
suggesting that pressing hard with the roller does not alleviate these problems. Large-scale
unevenness is also observed in the 0-degree tensile test specimens, where the valleys are once again
parallel to the fiber direction. Large-scale unevenness is also seen to some extent on the top-surface
of the 3 point bending specimen as can be seen on Figure 63.

Figure 63: Large-scale surface unevenness visible on a 3PB specimen

Because the depth of this large-scale unevenness is so close to the layer thickness, the whole top
layer was sanded off of some tensile test specimens to level the surface, exposing the interface
between the two topmost layers, which is more porous than the layers themselves.

5.2 Material properties


In this section, the material properties derived from tests are presented and compared to the values
taken from data sheets for Gurit SE84LV-HEC unidirectional prepreg and Gurit Corecell M80 and
M130 foam.
The material properties from the data sheets are given in Table 8 and Table 9 for the prepreg, the
foams respectively.
Fiber Tensile Tensile Compressive Compressive Interlaminar
volume strength modulus strength modulus shear
fraction strength
0° 55% 2458 MPa 134 GPa 1354 MPa 121 GPa 86,6 MPa
90° - 39,2 MPa 8,3 GPa - - -
Table 8: Material properties of SE84LV-HEC from data sheet

In the data sheets, the compressive properties in the transverse direction are missing and so are the
in-plane shear properties. In [28], tests have been performed to determine the in-plane shear
strength through the double-notch test method, which is found to be 58,9 [𝑀𝑃𝑎] with a standard

Page 86 af 125
cure in an oven. As the interlaminar shear strength in Table 8 are found from short beam bending
tests under a homogeneity assumption using the theory for long slender beams, this cannot be used
for design [2]. Instead, the in-plane shear strength from [28] is assumed to also be the through-
thickness shear strength.
Tensile Tensile Compressive Compressive Shear Shear Shear
strength modulus strength modulus strength modulus elongation
at break
Corecell 1,62 72 GPa 1,02 MPa 71 GPa 1,09 29 GPa 58%
M80 MPa MPa
Corecell 2,85 176 GPa 2,31 MPa 170 GPa 1,98 59 GPa 43%
M130 MPa MPa
Table 9: Material properties for Gurit Corecell M80 and M130 from data sheet

The material properties derived from tests are given in Table 10 and Table 11.
Fiber volume Tensile Tensile Compressive Compressive
fraction strength modulus strength modulus

0° 52% - 60% 1991 MPa 147,72 GPa - -


90° - 34,25 MPa 8,77 GPa 103,7 MPa 8,412 GPa
Table 10: Material properties of Gurit SE84LV-HEC derived from tests

The shear moduli in Table 11 are calculated from the tensile modulus, assuming isotropy using the
poissons ratio calculated from the tensile and shear moduli of Table 9. The strengths for the foam
are the true stress, assuming constant volume.
Tensile Yield 0,25% proof Ultimate Shear
Modulus strength strength strength modulus
Corecell M80 92,19 GPa 1,252 MPa 1,974 MPa 2,652 MPa 37,12 GPa

Corecell 176,05 GPa 1,826 MPa 3,759 MPa 4,705 MPa 59,02 GPa
M130
Table 11: Material properties of Gurit Corecell M80 and M130 derived from tests

5.2.1 Fiber and void volume fraction of SE48LV-HEC


The expected result was that the specimen with the unperforated release film had a lower fiber
volume ratio as excess resin would not be absorbed and also that the 3 layer specimen could have a
lower fiber volume ratio and could have more voids than the single layer specimen. The calculated
means and 95% confidence intervals are given in Table 12 where it is seen that both of the 1-layer
groups were found to have lower fiber content and higher void content than the 3-layer group of
specimen.

Page 87 af 125
3 layer specimen 1 layer 1 layer
Perforated release film
Perforated release film Unperforated release film

𝑣𝑓 0,603 0,555 0,537


mean
𝑐𝑜𝑛𝑓95% = [0,543; 0,663] 𝑐𝑜𝑛𝑓95% = [0,488; 0,622] 𝑐𝑜𝑛𝑓95% = [0,423; 0,651]

𝑣𝑣 0,034 0,055 0,045


mean
𝑐𝑜𝑛𝑓95% = [−0,020; 0,088] 𝑐𝑜𝑛𝑓95% = [−0,047; 0,158] 𝑐𝑜𝑛𝑓95% = [−0,058; 0,148]
Table 12: Fiber and void volume fraction of SE84LV-HEC

By plotting the void volume fraction against the fiber volume fraction on Figure 64, it is noted that
the 3-layer specimen are much closer grouped than the 1-layer specimen and also, there is an outlier
among the 1-layer group with unperforated release film.
The main possible errors in the method of the burn-off test are incomplete burning of matrix,
excessive burning of fiber, incorrect volume calculation due to nonnegligible amount of nail polish,
incorrect volume calculation due to unnoticed nonnegligible air bubbles stuck to the surfaces of the
specimen and discrepancies between densities of the actual fiber and matrix and the values listed in
the data sheets. While all of the specimen are found to be fibrous after burning, indicating that the
resin is properly burned off, it is noted that many of the 3 layer specimen has soot left on the sides
of the pot while none of the 1 layer specimen has. This suggests that the 3 layer specimen could
have unburned resin included in the weighing and/or that the 1 layer specimen could have excessive
fiber burn off.

Page 88 af 125
Figure 64: Void volume fraction plotted against fiber volume fraction

To investigate further into the possible errors, the weight of the nail polish and the calculated
densities are plotted against the fiber volume fraction in Figure 65 and Figure 66. It is seen that the
weight of the added nail polish are consistent among all specimen while the calculated densities of
the 1-layer specimen are much more spread out than for the 3-layer specimen and that the outlier
among the 1-layer specimen is also an outlier in terms of calculated density.

Figure 65: Nail polish weight plotted against fiber volume Figure 66: Specimen density plotted against fiber volume
fraction fraction

Page 89 af 125
As multiple of the specimen within each group are cut from the same cured laminate strips, the lack
of clustering among the calculated densities of the 1-layer specimen indicates that errors exist in the
weighing of the specimen. This error may then have to do with the ratio of surface to volume as
submerging the specimen into water has been difficult due to surface tension and some specimen
has had to be submerged multiple times to get rid of air bubbles clinging to the surface.
To test if the thickness of lamination or the choice of perforated versus unperforated release film
can be said to influence the fiber and void volume fractions despite the errors, hypothesis testing is
applied by testing if the means of two dataset are equal or if the hypothesis can be rejected at a 5%
significance level. In Table 13, the result of these hypothesis test are given where the difference of
means are given when the hypothesis is rejected and the p-value are given when the hypothesis is
accepted. The outlier of the group with unperforated release film has been excluded from this
hypothesis testing but the conclusions are the same, just with slightly different p-values. As is seen,
the impact of laminate thickness is not what would be expected if more voids exist in the
interlaminar regions. Also, the choice of release film cannot be proven to impact the fiber and void
volumes, although these p-values are very low, casting doubt on the conclusions.
3 layer specimen 3 layer specimen 1 layer
Perforated release film
Perforated release film Perforated release film
&
& &
1 layer
1 layer 1 layer
Perforated release film Unperforated release film
Unperforated release film
𝑣𝑓 being equal Rejected Rejected Accepted
𝑥̅ − 𝑦̅ = [0,04 ; 0,056] 𝑥̅ − 𝑦̅ = [0,045 ; 0,076] 𝑝 = 0,108

𝑣𝑣 being equal Rejected Accepted Accepted


𝑥̅ − 𝑦̅ 𝑝 = 0,214 𝑝 = 0,112
= [−0,037 ; −0,005]
Table 13: Hypothesis testing of fiber and void volume fractions

As it is not known for sure which errors exist in these tests, the exact fiber volume fraction or void
content cannot be determined uniquely. An approximate upper bound on the fiber volume fraction
and void ratio of the 3 layer specimen can be established from the data as a lower weight after
burning would reduce both of these calculated values. A lower bound on the hypothesizing that the
3 layer specimen had no voids and reducing the after burn weight to that which would result in no
calculated voids, a lower bound of the fiber volume fraction can be placed in the interval 0,52-0,54
which matches nicely with the 55% fiber volume stated in the data sheet as the layup is unlikely to
be completely free of voids.
5.2.2 Composite tensile test – longitudinal direction
The tensile tests in the longitudinal direction are tested until failure with the stress-strain diagram
shown in Figure 67. When individual fibers fractures during loading, it happens with such violence
that the extensiometer is seen to jump which leads to discontinuities in the strain data. While the
data could have been cleaned up by assuming continuity across these jumps, no attempt to do so

Page 90 af 125
have been conducted in this project, which leads to only the ultimate stress and the stiffness before
the discontinuities being reliable data from these tests. The failure strain is not noted as it would
rely heavily on assumptions and interpretation of the data.

Figure 67: Stress-strain diagram of SE84LV-HEC, longitudinal direction

The stiffness data is very noisy, possibly due to an overly slow testing speed. The stiffness plot
shown on Figure 68, where the stiffness is calculated as an average over 100 datapoints indicate a
possible strain stiffening effect which is not accounted for. In this plot, instantaneous jumps in
stiffness, as seen by the vertical lines that extend beyond the range of the plot, are related to fiber
fracture. It is seen that the specimen are free of damage in the first 0,0045 % of strain.

Page 91 af 125
Figure 68: Stiffness plot of SE84LV-HEC, longitudinal direction

The stiffness of the specimen are calculated per iso 527, where the stiffness is assumed to be the
slope of a straight line connecting the stress at 0,0005 % strain and 0,0025% strain. The stiffness
and strength of all specimen and the mean values are given in Table 14 along with the 95%
confidence interval of the stiffnesses and of the mean stiffness and strength. The stiffness is found
to be higher than that stated in the data sheet while the strength is found to be lower.
Specimen Stiffness per iso Ultimate strength
no standard [GPa] [Mpa]
1 155,29 +/- 1,08 2024
2 153,95 +/- 2,88 1960
3 155,34 +/- 0,93 1896
4 150,88 +/- 0,26 1944
5 145,31 +/- 0,32 2056
6 145,83 +/- 0,61 1801
7 141,54 +/- 0,31 2066
8 142,20 +/- 0,15 2104
9 146,42 +/- 0,54 1966
10 143,97 +/- 0,34 1825
11 155,00 +/- 0,54 1980

Page 92 af 125
12 146,99 +/- 0,24 2125
13 136,06 +/- 0,22 1963
14 148,02 +/- 0,21 1989
15 150,00 +/- 0,63 2167
Mean and 147,72 +/- 3,17 1991 +/- 57
95th
percentile
interval
Table 14: Stiffness and strength of SE84LV-HEC, longitudinal direction

Upon failure, these specimen exploded.


5.2.3 Composite tensile test – transverse direction
Like the longitudinal specimen, the stiffness is calculated by a straight line connecting the stress at
0,0005 % strain and 0,0025% strain. The stiffness and strength is given in Table 15. These
specimen failed unspectacularly by a clean fracture across the cross-section.
Specimen Stiffness per iso Ultimate strength
no standard [GPa] [Mpa]
1 7,05 +/- 1,48 37,0
2 7,29 +/- 0,88 30,4
3 8,64 +/- 0,03 35,7
4 8,92 +/- 0,03 40,2
5 8,79 +/- 0,03 26,1
6 8,67 +/- 0,03 33,3
7 8,90 +/- 0,03 38,7
8 8,75 +/- 0,03 33,7
9 8,72 +/- 0,03 33,1
Combined 8,77 +/- 0,10 34,25 +/- 3,33
mean and
95th
percentile
interval
Table 15: Stiffness and strength of SE84LV-HEC, transverse direction

5.2.4 Composite compressive test – longitudinal direction


All of these specimen failed by splitting/crushing at the loaded end as shown in Figure 69 and with
much lower stiffnesses and strengths than given by the data sheet. For this reason, all of these

Page 93 af 125
specimens are considered invalid and it can be concluded that tabs and strain gages must be used to
accurately conduct compressive test of composites in the longitudinal direction.

Figure 69: End-failure of SE84LV-HEC compressive specimen

5.2.5 Composite compressive test – transverse direction


Only specimen 5 failed near the loaded end while all others failed in the section that were supported
by the testing fixture. The stiffness and strength of specimen 5 are in agreement with the majority of
the test specimen and are therefore considered valid. The stress-strain diagram and stiffness diagram
are shown in Figure 70 and Figure 71 respectively. As the stiffness is not constant, the method of
determining stiffness through statistical search of an interval with approximately constant stiffness
is used as it is for the 3-point bending tests. The results are given in Table 16.

Page 94 af 125
Figure 70: Stress-strain diagram of SE84LV-HEC in compression, transverse direction

Figure 71: Stiffness diagram of SE84LV-HEC in compression, transverse direction

Page 95 af 125
Specimen Stiffness per statistical Ultimate strength
no method [GPa] [Mpa]
(strain interval in
parentheses)
1 7,871 +/- 0,0,049 102,8
(0,0048 – 0,0081)
2 8,003 +/- 0,043 73,1
(0,0048 – 0,0099)
3 8,711 +/- 0,038 111,3
(0,0047 – 0,0112)
4 8,744 +/- 0,039 128,9
(0,0046 – 0,0126)
5 8,731 +/- 0,040 102,4
(0,0048 – 0,0112)
Combined 8,412 +/- 0,542 103,7 +/- 25,1
mean and
95th
percentile
interval
Table 16: Stiffness and strength of SE84LV-HEC in compression, transverse direction

Inspection of the specimen after failure shows a fracture at an angle to the loaded direction as
shown in Figure 72.

Figure 72: Fracture of compression specimen under a microscope

5.2.6 Gurit Corecell M80 tensile test


All specimen yielded but failed without necking by fracturing straight across the cross-section. The
stress-strain diagram and stiffness diagram from the engineering stress/strain are given in Figure 73
and Figure 74 respectively. The stiffness is seen to drop off immediately from the start of the test,
indicating that the cross-sectional area change quickly becomes unneglectable before yielding starts

Page 96 af 125
to take effect, thus making it difficult to identify a region of approximately constant stiffness. By
applying the constant volume assumption to plot the true stress against the logarithmic strain, the
stiffness is initially approximately constant for a longer strain interval as seen by the zoomed-in
view of Figure 75 and Figure 76.

Figure 73: Engineering stress-strain diagram for Corecell Figure 74: Stiffness diagram for Corecell M80 tensile test,
M80 tensile test engineering stress/strain

Figure 75: Zoomed in view of the stiffness diagram of M80 Figure 76: Zoomed in view of the stiffness diagram of M80
from engineering stress/strain from true stress/strain

The stiffness is then calculated from the true stress/strain as this stiffness should be equal to the
stiffness with engineering stress/strain at infinitesimal strain levels. The calculated stiffnesses and
strengths are presented in Table 17. To calculate the stiffness, the statistical method is used to
identify the end of the linear region. As this method will place the end of the linear region very
early, using the end of the linear region as the yield strength is not usable for building a bilinear
hardening model. Instead, the stress at which there are 0,25% plastic strain are used.
Twelve specimen has been tested but the first is invalidated due to errors with the extensiometer.

Page 97 af 125
Stiffness per Yield strength, 0,25% Proof Ultimate
statistical method end of linear strength strength
[GPa] region
[MPa] True stress
(strain interval in [MPa]
[MPa]
parentheses)
Specimen 2 89,20 +/-1,07 1,251 1,906 2,551
(0 – 0,0140)
Specimen 3 92,80 +/-1,41 1,420 2,007 2,633
(0 – 0,0152)
Specimen 4 90,71 +/-1,13 1,203 1,944 2,615
(0 – 0,0132)
Specimen 5 92,36 +/-1,23 1,115 1,913 2,530
(0 – 0,0120)
Specimen 6 88,48 +/-1,23 1,480 1,993 2,674
(0 – 0,0166)
Specimen 7 94,87 +/-1,19 1,124 1,975 2,634
(0 – 0,0119)
Specimen 8 92,80 +/- 1,20 1,485 2,034 2,743
(0 – 0,0160)
Specimen 9 96,59 +/- 1,20 0,828 1,969 2,712
(0 – 0,0085)
Specimen 10 94,70 +/- 1,20 1,212 1,976 2,731
(0 – 0,0128)
Specimen 11 91,14 +/- 1,20 1,427 2,005 2,654
(0 – 0,0156)
Specimen 12 90,41 +/- 1,17 1,224 1,990 2,696
(0,0037 – 0,0134)

Mean 92,19 +/- 1,68 1,252 +/-0,132 1,974 +/-0,027 2,652 +/-0,047
Table 17: Stiffnesses and strengths of Gurit Corecell M80 foam

Assuming the material to be isotropic, a poissons ratio of 0,241 are calculated from the data sheet
data which are then used to calculate a shear stiffness of 37,12 MPa using the mean stiffness from
these tests.
To form the bilinear model, the tangent modulus are now to be found. One method could be to draw
a straight line from the yield stress to the highest true stress for each curve in Figure 77 but this
would make the calculated tangent modulus stiffer for the specimen that failed earlier. Instead,

Page 98 af 125
Ludwik’s model in equation (3-131) is fitted to the stress versus plastic strain shown in Figure 78 in
order to extrapolate the true stress-strain curve so that the end-point strain of the tangent modulus is
the same for all specimen. For each specimen, (3-131) is fitted to an interval starting at the 0,25%
proof stress and ending at the highest stress value. This model is generally not a perfect fit as seen
on Figure 79 which may in part be explained by the stress/plastic strain graph swinging slightly into
negative strains due to the elastic modulus being calculated lower than the actual modulus at
infinitesimal strains.

Figure 77: True stress versus logarithmic strain, M80 tensile Figure 78: True stress versus logarithmic plastic strain, M80
test tensile test

Figure 79: Fit of Ludwik's model onto specimen 2 of M80 tensile test

The coefficients for Ludwik’s model and the calculated tangent moduli 𝐾𝑡 are given in Table 18.
The end point for calculating the tangent modulus is chosen to be at 0,7 strain.

Page 99 af 125
Specimen 2 3 4 5 6 7 8 9 10 11 12
no.
𝒦 3,426 3,529 3,466 3,536 3,421 3,622 3,521 3,323 3,579 3,493 3,528
𝓃 0,094 0,091 0,092 0,098 0,087 0,096 0,088 0,096 0,094 0,089 0,091

𝜎𝑌 , 𝑀𝑃𝑎 0 0 0 0 0 0 0 0 0 0 0
𝐾𝑡 , 𝑀𝑃𝑎 8,726 8,833 8,877 9,236 8,354 9,365 8,676 9,381 9,141 8,649 8,837
Table 18: Fitting coefficients for Ludwik's model and bilinear tangent modulus for M80 tensile test

The mean tangent modulus is 8,9059 MPa.


To test whether this bilinear model is suitable, 3 point bending tests have been conducted on the
M80 foam without any composite skins and these tests are now compared to a finite element model
using solid-shell elements, with displacement boundary conditions at the supports and a flexible
remote displacement as the load applicator. The load-displacement graph and the stiffness graph are
shown on Figure 80 and Figure 81 respectively and it is seen that the initial stiffness is fairly well
modeled but that using the 0,25% proof strength as the yield strength does not capture the onset of
plastic deformation properly. Using the stress at the end of the linear region as the yield strength
places the yield at a more appropriate stress so the drop in stiffness follows what is seen in
experimental data but the stiffness between 35 and 50 mm displacement still levels out at a too high
stiffness, indicating that the tangent modulus must also be lowered.

Figure 80: Comparison of exprimental and simulated data of Figure 81: Comparison of exprimental and simulated data of
3PB of M80 foam using bilinear isotropic hardening with 3PB of M80 foam using bilinear isotropic hardening with
experimentally derived material data, Load-Displacement graph experimentally derived material data, Stiffness graph

To get the bilinear model to yield the same force at the end of 50 mm displacement as the
experimental results, the tangent modulus must be set to practically zero but with 10%
overprediction of the load at 50 mm deflection, a tangent modulus of 6 can be used.
To make a multilinear model, stress levels are taken from the graphs in Figure 78 until the
maximum load, after which the extrapolated values are taken. The mean values are given in Table
19.

Page 100 af 125


Plastic 0 0,00075 0,001 0,002 0,004 0,008 0,014 0,04 0,1
strain
level
Stress, 1,0575 1,6713 1,7356 1,9142 2,0953 2,2651 2,3912 2,6027 2,7888
MPa
Table 19: Stress levels for multilinear model of M80 foam

Within this project, attempts at applying this multilinear model to the 3-point bending of a 25mm
thick M80 foam plate has been unsuccessful as ANSYS fails to converge due to distorted elements,
even when the mesh is very fine.

The following in this subsection is added in V2.0 of this report, the 11th of august 2020
An error in the data entry of ANSYS has been located, as the data in Table 19 was entered as-is in
megapascals while ANSYS per default assumes the stresses to be Pascals. When this is corrected,
the simulation does not fail to converge but using the experimentally derived values for stress
versus plastic strain, the simulation with multilinear isotropic hardening fails to capture the
nonlinear behavior properly as seen on Figure 82. Multiple simulated results are shown, which are
different refinement levels of mesh and timestepping, which does not suggest that the discrepancy is
due to numerical errors. Lowering the stress level at zero plastic strain has also been attempted but
yields practically no difference to the results. To control that the issue is not in the symmetry
boundary conditions, a simulation of a full plate has also been conducted, which gives similar
results to the quarter model simulations.

Figure 82: Load-displacement graph of 3PB of M80 foam. Comparison of experimental data and simulated with multilinear
plasticity model

Page 101 af 125


5.2.7 Gurit Corecell M130 tensile test
All specimen yielded but failed without necking by fracturing straight across the cross-section. Like
the M80 specimen, the stiffness drops off immediately after testing begins and the search for the
approximately linear region using the statistical method is done on the true stress versus logarithmic
strain.

Figure 83: True stress-strain diagram of M130 tensile tests Figure 84: Stiffness diagram of M130 tensile test, from true
stress and logarithmic strain

Stiffness per Yield strength, 0,25% Proof Ultimate strength


statistical method end of linear strength
True stress
[GPa] region
[MPa]
[MPa]
(strain interval in [MPa]
parentheses)
Specimen 1 173,21 +/- 1,796 2,0383 3,8762 4,7086
(0 – 0,0117)
Specimen 2 176,15 +/- 1,265 1,8084 3,7613 4,6932
(0 – 0,0102)
Specimen 3 182,54 +/- 1,490 2,1207 3,8709 4,8195
(0,0016 – 0,0115)
Specimen 4 174,36 +/- 1,436 1,6165 3,6759 4,6870
(0 – 0,0093)
Specimen 5 166,82 +/- 1,313 2,2083 3,7560 4,6900
(0,0059 – 0,0128)

Page 102 af 125


Specimen 6 178,55 +/- 1,548 1,8482 3,7862 4,7556
(0 – 0,0120)
Specimen 7 179,60 +/- 1,356 1,5620 3,7288 4,6623
(0 – 0,0086)
Specimen 8 174,30 +/- 1,203 1,9126 3,6241 4,6536
(0 – 0,0110)
Specimen 9 178,88 +/- 1,424 1,3203 3,7546 4,6743
(0 – 0,0073)
Mean 176,05 +/- 3,530 1,826 +/-0,220 3,759 +/-0,063 4,705 +/-0,040
Table 20: Stiffness and strength of M130 tensile test

The coefficients for Ludwik’s model and the calculated tangent moduli 𝐾𝑡 are given in Table 21.
The end point for calculating the tangent modulus is chosen to be at 0,7 strain.
Specimen 1 2 3 4 5 6 7 8 9
no.
𝒦 5,6960 5,8092 6,0069 5,8361 5,7203 5,8888 5,9094 5,9178 5,7472
𝓃 0,0616 0,0670 0,0682 0,0711 0,0656 0,0682 0,0707 0,0754 0,0659

𝜎𝑌 , 𝑀𝑃𝑎 0 0 0 0 0 0 0 0 0
𝐾𝑡 , 𝑀𝑃𝑎 11,299 12,208 12,778 12,737 11,816 12,516 12,866 13,396 11,962
Table 21: Coefficients of Ludwik's model and the calculated tangent modulus for M130 foam tensile tests

The mean tangent modulus is 12,398 MPa.


The stress values corresponding to different levels of plastic strain to be used for a multilinear
hardening model are given in Table 22.
Plastic 0 0,00075 0,001 0,002 0,004 0,008 0,014 0,04 0,1
strain
level
Stress, 2,637 3,240 3,353 3,656 3,967 4,227 4,400 4,667 4,913
MPa
Table 22: Stress and strain values for a multilinear hardening model, M130 tensile tests

The following in this subsection is added in V2.0 of this report, the 11th of august 2020
As with the M80 foam, a data entry error lead to nonconvergence of multilinear plasticity models in
ANSYS using the data in Table 22. With this error corrected, the simulation does not fail to
converge but the load-carrying capability of the nonlinear region is overestimated as shown on
Figure 85.

Page 103 af 125


Figure 85: Load-Displacement graph of 3PB of M130 foam. Comparison of experimental results to simulated with multilinear
plasticity.

5.3 3-point bending


All of the specimen are seen to exhibit unneglectable compression of the core at the contact with the
central load applicator as seen on Figure 86, followed by core yielding and then failure of the top
skins. Apart from specimen 11 and 12, all specimen failed by a through-thickness crack of the top
skin. Specimen 11 and 12 has had a vacuum failure during curing so that the vacuum pressure was
up to 250 mBar and these specimens failed by delamination.

Figure 86: Compression of core seen at the middle of a 3PB test specimen

During the testing of specimen 11, delamination occurred between the core and the first layer of
composite while specimen 12 had a delamination of only the top layer of the top skin with the
remaining layers attached to the core.

Page 104 af 125


For a description of the typical 3 point bending of these laminates, the load-displacement graph and
the stiffness graph of specimen 1 are shown in Figure 87 and Figure 88 respectively. The first
millimeter of displacement is dominated by contact nonlinearities as the load applicator come in
contact with the whole width of the test specimen. From 1 to 4 millimeters, the stiffness is mostly
stable but decreasing as the top skin is deformed to the shape of the load applicator while the core
deforms elastically. Between 4 and 5 mm, there’s a dip in the stiffness, the cause of which is not
known, but it is accompanied by a ticking sound, so it may be slight delamination or growth of
internal fractures in the laminate. From 5 to 8 mm, there’s a transition region between elastic and
plastic domination, followed by a region of near constant load until 15mm where the deformation of
the cross-section becomes so severe that the stiffness is dropping until the top skin fails.

Figure 87: Load-displacement graph of specimen 1 Figure 88: Stiffness graph of specimen 1

In the following, the load-displacement and stiffness graphs for all specimen are shown, grouped
according to laminate. All of the laminates with M80 core deform beyond the rules required test
range of 12,7 mm before failing and have a yield-like behavior of the laminate as a whole. The
laminates with M130 core fail much earlier and with less yield-like behavior. The strength and
stiffness of having a skin thickness of 4 layers and a core of M130 foam are near those of an M80
core with 8 layers for each skin but the weight of the former are around 990g while the weight of
the latter are around 1300g, showing the importance of choosing the core material.

Page 105 af 125


Figure 89: 3PB load-displacement curve, 4 Figure 90: 3PB stiffness graph, 4 layer skins,
layer skins, 25mm M80 core 25mm M80 core

Page 106 af 125


Figure 91:3PB load-displacement curve, 4 layer skins, 25mm Figure 92:3PB stiffness graph, 4 layer skins, 25mm M130 core
M130 core

Figure 93:3PB load-displacement curve, 4 layer skins, 10 mm Figure 94:3PB stiffness graph, 4 layer skins, 10 mm M80 core
M80 core

Page 107 af 125


Figure 95: 3PB load-displacement graph, 8 layer skins, 25mm Figure 96: 3PB stiffness graph, 8 layer skins, 25mm M80 core
M80 core

Figure 97: 3PB Load-displacement graph, 14 layer skins,


25mm M80 core Figure 98: 3PB stiffness graph, 14 layer skins, 25mm M80 core

When simulating the bending of these plates using a model with a solid core in ANSYS to capture
the nonlinear effects, the stiffness of the specimen is seen to rise before yielding, like it was seen for
the steel plate. As the stiffness is observed to drop in the experimental values, it indicates that the
nonlinearity of the core material must be simulated to start very early, which cannot be done with a
bilinear hardening model. Like for bending of the core material by itself, employing a multilinear
hardening model was not successful for the sandwich plates either. The load-displacement curve
and the stiffness curves seen in Figure 99 and Figure 100 compare the simulated results with the
stiffness from the data sheet to the simulated result with the tested stiffness and to the experimental
result of bending specimen 1. The flexibility of the testing rig has been added to the simulated data
to compare against the experimental data. The importance of modelling plastic behavior can be seen
very clearly as the maximum load of the simulated data is sensitive to the yield strength but that the
curve cannot be made to follow completely by just a bilinear model.
The material for the core that best fit the experimental results of specimen 1 are a yield strength of
0,95 MPa and a tangent modulus of 2 MPa but this vastly underestimate the strength of specimen 7
which has 8 layer thick skins.

Page 108 af 125


Figure 99: Load-Displacement curve comparison between Figure 100: Stiffness graph comparison between simulated and
simualted and tested data, 3PB, specimen 1 tested data, 3PB, specimen 1

In the following, the stiffness of the test specimen are plotted along with the simulated data in
Figure 102 to Figure 113. The legend is the same for all of these figures and are given in Figure
101.

Figure 101: Legend for comparison graphs.

Page 109 af 125


Figure 102: Laminate stiffness versus Figure 103: Laminate stiffness versus Figure 104: Laminate stiffness versus
skin thickness core thickness core density. The densities here are
considered the identification of the
foams in the Gurit Corecell M-series

Figure 105: Energy absorption versus Figure 106: Energy absorption versus Figure 107: Energy absorption versus
skin thickness core thickness core density

Page 110 af 125


Figure 108: Yield load versus skin Figure 109: Yield load versus core Figure 110: Yield load versus core
thickness thickness density

Figure 111: Maximum load versus skin Figure 112: Maximum load versus core Figure 113: Maximum load versus core
thickness thickness density

In order to use the computational methods to search for optimum layups in a preliminary design
phase prior to conducting tests, it is necessary to know the percent of error of the simulations
relative to the experiments but also how much the individual tests differ from the mean values. As
there are too few experimental results for each layup to give a meaningful 95th percentile confidence

Page 111 af 125


interval, the spread of the experimental data is given in Table 23 as a mean and a spread where the
spread is the percentwise difference from the mean to the datapoint furthest from the mean. The
data is generally grouped pretty closely, indicating that defects in the layup are pretty consistent and
that the use of chemical release agent versus release film on the glass plate does not have a large
impact – if any – on these figures and the same for the use of perforated versus unperforated release
film. It is also worth noting that the spread in the strengths and energy absorptions of the specimen
with M130 core are larger than in the specimen with the M80 core as the failure behavior of the
beam as a whole is closer to the brittle failure of the carbon fiber composite when the core is stiffer
and doesn’t show as much yielding before fiber failure.
4-layer skins 8-layer skins 14-layer skins 4-layer skins 4-layer skins
25mm core 25mm core 25mm core 25mm core 10 mm core
M80 foam M80 foam M80 foam M130 foam M80 foam
Stiffness, 1502,7 2039,2 2720,7 2098,6 531,6
N/mm +/- 2,1% +/- 2,3% +/-1,3% +/-1,4% +/-0,7%
Energy 72,93 115,48 151,88 88,96 38,04
absorption, J +/- 3,1% +/-1,7% +/-0,4% +/-8,9% +/-3,3%
Yield load, N 4036,6 6868,3 9096,7 9521,7 3947,7
+/- 37% +/-4,6% +/-3,8% +/-6,0% +/-6,8%
Highest 7703,1 14334 19061 14439 4889,0
measured +/-3,1% +/-2,5% +/-0,9% +/-9,0% +/-5,3%
load, N
Table 23: Summary of experimental data, 3-point bending tests

The yield load data has a higher spread than the other data for each layup which is caused by the
way that this is calculated as the end of the linear region determined by the statistical method. This
method is more sensitive than if it was determined by other methods such as a fixed ratio in drop of
stiffness or a fixed percent of nonlinear displacement which are not chosen as the fixed values
would be chosen somewhat arbitrary and could cause either a system that is very sensitive to noise
or where the range of linear displacement is overestimated, leading to lower calculated stiffnesses
and higher calculated yield loads.
The high spread in the calculated yield load of the specimen with 4 layer skins and a 25mm M80
core are due to specimen 11 having a much earlier calculated end of the linear region than the
others. By examining the stiffness graph in Figure 90, it is seen the stiffness graph follows the other
two specimen pretty well beyond the calculated yield load and thus it is assumed that it is an artifact
of the chosen method as discussed above. If the yield load of specimen 11 is considered an outlier
and are omitted, the mean yield load of that layup are 4788,3 N/mm with a spread of +/-4,4%. As
the spread with specimen 11 omitted is very similar to that of the other layups, this mean yield load
is chosen instead for the evaluation of the simulation errors.
The simulation error are shown in Table 24, Table 25, Table 26 and Table 27 for the stiffness,
energy absorption, yield load and maximum load respectively. From these tables and the figures

Page 112 af 125


above, it can be seen that the 1D beam simulations and the ANSYS shell model are close to each
other in terms of determining the stiffness and that both overestimate the stiffness due to not
including compression of the core but that they approach the correct result as the skins get stiffer.
This is due to the stiffer skins being less able to produce a small radius of curvature by themselves
and thus they do not load put as concentrated a load on the core where the loads and supports are,
which lessens the influence of core normal compression.
It is also seen that the simulations with the datasheet core data of M80 foam are generally closer in
stiffness to the experimental results than with the tested data for all but the full model. It is not
completely clear whether this is due to the apparent decreasing stiffness of the core material during
the approximately linear deformation or if it is a coincidence that the difference in datasheet and
tested values are approximately as much as the models overestimate the stiffness. The latter is
assumed as sophisticated nonlinear material models are not applied to test the former and since the
full model matches relatively well when the tested stiffness is used.

Stiffness error 4-layer skins 8-layer skins 14-layer 4-layer skins 4-layer skins
skins
25mm core 25mm core 25mm core 10 mm core
M80 foam M80 foam 25mm core M130 foam M80 foam
M80 foam
1D beam, tested 35,9% 27,8% 18,3% 19,3% 22,0%
data
1D beam, 13,4% 3,7% -5,3% 19,3% 7,3%
datasheet data
2D beam, tested 26,6% 25,4% 27,4% 11,5% 21,1%
data
2D beam, 6,3% 3,9% 6,2% 11,5% 7,2%
datasheet data
Shell model, 38,2% 30,3% 20,7% 18,5% 23,1%
tested data
Shell model, 15,5% 5,9% -3,3% 18,5% 8,4%
datasheet data
Full model, 0,3% -8,2% 12,9% -0,2% -44,9%
tested data 16,1% 15,3% 10,5% 4,4% 20,4%
Full model, -16,4% -11,0% -6,3% 0,3% -22,7%
datasheet data -1,2% -3,2% -6,7% 4,4% 6,9%
Table 24: Simulation error, stiffness

The relatively large error in the stiffness full model with 10 mm core could be due to a meshing
issue as the discretization through the thickness is set to a certain number of elements while the face

Page 113 af 125


size is set to a certain dimension, which makes the aspect ratio of the core elements worse for the 10
mm thick model than for the 25mm thick models as these numbers have been equal for all
simulations. This might also be the cause of relatively large errors in the max load, yield load and
energy absorption.
The following paragraph is added in V2.1 the 17th of august 2020.
The large stiffness error of the full model simulation of the specimen with 10mm core is found to be
somewhat related to the mesh as refining it to have a face size of 2mm so the aspect ratio is lowered
from 4,1 to 0,9 makes it overpredict the stiffness by 11,5% instead of underpredicting by almost
45%. This does however come at the cost of increasing computational time from 11 minutes to 4
hours.
As mentioned in the addition related to multilinear plasticity in the end of this section, the contact
between the skin and the core creates an error which is very mesh dependent when the default
augmented lagrange formulation is used. When the pure lagrange method is used with the original
mesh, the computation time is 21 minutes and the stiffness during the first millimeter is
overpredicted by 24% instead of an underprediction by almost 45%.
With the pure lagrange contact formulation, the weird rise in stiffness is also gone and the stiffness
is dropping at a steady rate until yielding sets in, which it does gradually and thus the statistical
method of determining the stiffness is applied. This leads to the yield load being calculated much
too low as the end of the “linear” region is before any plastic strain occurs. The newly calculated
values have been inserted into the result tables, Table 24, Table 25, Table 26 and Table 27, but not
in the figures, Figure 102 through Figure 113.
End of addition in V2.1
The error of energy absorption, yield load and maximum load are very dependent on the modeling
of the very nonlinear behaviors of the structure and the materials and thus, very large errors are seen
for all but the full model, which fits the first three layups relatively well. The max load error is
small for some of the simulations of beam elements and of the 2D core but as it isn’t consistent with
increasing skin stiffness, it shouldn’t be interpreted as a result of good modeling.
Energy 4 layer skins 8 layer skins 14 layer skins 4 layer skins 4 layer skins
absorption error
25mm core 25mm core 25mm core 25mm core 10 mm core
M80 foam M80 foam M80 foam M130 foam M80 foam
1D beam, tested -20,9% -49,9% -59,3% 28,7% -24,0%
data
1D beam, -14,3% -47,4% -58,7% 28,7% -17,1%
datasheet data
2D beam, tested -55,2% -60,3% -58,6% -42,8% -23,7%
data
2D beam, -53,6% -56,6% -56,6% -42,8% -17,0%
datasheet data

Page 114 af 125


Shell model, 127% 85,3% 74,5% 148,4% 42,0%
tested data
Shell model, 90,2% 50,7% -20,3% 148,4% 25,1%
datasheet data
Full model, 6,7% -22,3% -9,0% 55,9% -39,6%
tested data 18,5% -9,9% -6,5% 66,8% -4,7%
Full model, 2,0% -17,3% -12,2% 55,1% -16,4%
datasheet data 13,4% -13,2% -10,0% 66,8% -8,3%
Table 25: Simulation error, energy absorption

Yield load error 4 layer skins 8 layer skins 14 layer skins 4 layer skins 4 layer skins
25mm core 25mm core 25mm core 25mm core 10 mm core
M80 foam M80 foam M80 foam M130 foam M80 foam
1D beam, tested 184% 108% 68,0% 151% 50,0%
data
1D beam, 184% 108% 68,0% 151% 50,0%
datasheet data
2D beam, tested 65,9% 63,3% 69,2% 32,1% 49,9%
data
2D beam, 74,9% 78,3% 78,0% 32,1% 50,0%
datasheet data
Shell model, 447% 391% 359% 265% 116%
tested data*
Shell model, 357% 299% 267% 265% 90,9%
datasheet data*
Full model, 7,7% -1,2% -7,5% 37,0% -40,2%
tested data -42,6% -51,7% -75,9% -73,2% -54,2%
Full model, 26,9% 13,1% -6,4% 22,3% -20,1%
datasheet data -58,4% -53,5% -73,5% -73,2% -68,8%
Table 26: Simulation error, Yield load

*The Shell model in ANSYS does not include core yielding and also fail to predict any the failure
of the skins, meaning that the deformation is approximately linear until failure.
Max load error 4 layer skins 8 layer skins 14 layer skins 4 layer skins 4 layer skins
25mm core 25mm core 25mm core 25mm core 10 mm core
M80 foam M80 foam M80 foam M130 foam M80 foam

Page 115 af 125


1D beam, tested 76,9% -0,5% -19,8% 65,3% 21,1%
data
1D beam, 77,0% -0,4% -19,8% 65,3% 21,1%
datasheet data
2D beam, tested 3,3% -21,8% -19,2% -12,9% 21,1%
data
2D beam, 9,0% -14,5% -15,0% -12,9% 21,1%
datasheet data
Shell model, 240,7% 135% 119% 140,6% 74,8%
tested data
Shell model, 184,7% 91,3% 75,2% 140,6% 54,2%
datasheet data
Full model, 3,0% -35,1% -21,3% 6,8% -41,8%
tested data 8,0% -29,9% -22,5% 9,4% -4,4%
Full model, 2,4% -29,9% -21,8% 9,2% -23,8%
datasheet data 7,2% -29,7% -23,3% 9,4% -5,0%
Table 27: Simulation error, maximum load

The shell model calculates a slightly larger stiffness than the beam model, which indicates that there
are some stiffening effect from prevention of anticlastic curvature. The stiffening effect is found by
dividing the stiffness of the shell model by the stiffness of the beam model and are given in Table
28 and while these are not generally applicable to all laminated structures and are only applicable to
these layups when the beam width is 275mm, the values are used to assess the magnitude of the
stiffening effect. The stiffening effect on these specimen are relatively small compared to the
overprediction of stiffness, the possible variance in the stiffness of the core material and compared
to the spread in the experimental stiffnesses. The stiffening effect of completely preventing
anticlastic curvature of a homogeneous isotropic beam with a Poisson’s ratio of 0,3 is 9,89% so
compared to this, the stiffening effect on these specimen is also considered small. It is unexpected
that the layup with the thinner core exhibits a smaller stiffening effect than the similar layup with a
thicker core since the opposite is the case for homogeneous isotropic beams as described in section
3.5. As no attempt is made to generally quantify this stiffening effect on laminated beams, this is
not further investigated. That the stiffening effect seems negative on the layup with M130 core is
assumed to be due to numerical errors in ANSYS.
4 layer skins 8 layer skins 14 layer skins 4 layer skins 4 layer skins
25mm core 25mm core 25mm core 25mm core 10 mm core
M80 foam M80 foam M80 foam M130 foam M80 foam
Stiffening 1,71% 1,98% 2,02% -0,66% 0,84%
effect of

Page 116 af 125


beam/plate
width
Table 28: Stiffening effect of laminated beam width

The shell model is not considered to bring significantly more value to the problem of 3 point
bending than a beam model and are thus not considered valuable for the optimization problem.

The following in this subsection is added in V2.0 of this report, the 11th of august 2020. Expanded in
V2.1, the 17th of August 2020.
With the multilinear isotropic hardening model for modeling plasticity of the core, the model does
not converge, even with very small timesteps and very fine meshes, when the contact between the
skin and core is set to “program controlled” which uses the augmented lagrange formulation. By
setting the contact formulation to pure lagrange, it does converge but similarly as for simulation of
the foam by itself, the nonlinear behavior is not simulated well and the stiffness after core yielding
is overpredicted significantly.
The stiffness prior to yielding is also seen to be much higher than what is measured and what is
simulated with the bilinear model. This is not found to be due to too coarse mesh or timestep size
and also persists when the material nonlinear effects are disabled for the core. The difference is
found to be mostly due to the bilinear models having been meshed with a 5mm face size and 8
through-thickness divisions of the core, giving an aspect ratio of 1,6. Lowering the number of
through-thickness divisions raises the calculated stiffness, as does refining the face size to 3,2mm to
make the aspect ratio close to 1 while there’s 8 through-thickness divisions.
It is found to be connected to the contact formulation between the skins and the core as having the
pure lagrange formulation is about 5% stiffer than the default augmented lagrange formulation
when the mesh is fine and the aspect ratio is close to 1. When the mesh is coarser and the aspect
ratio is close to 1, the difference is about 8% and when the mesh is as used in the simulations in this
report, the difference is 15%. The results with the pure lagrange formulations are in close agreement
with only 1% difference between the fine mesh with a good aspect ratio and the coarser mesh with a
worse aspect ratio.

Page 117 af 125


5.4 Perimeter shear strength
For the perimeter shear strength tests, the typical response of loading starts with an initial
nonlinearity as the test specimen conforms to the testing fixture, followed by a short interval of
approximately linearly elastic response before the core starts to yield. After some yielding, the
specimen may make some sounds that indicate that internal fractures grow, accompanied by dips in
the stiffness. The first major peak is often due to the first or second layer from the top splitting
parallel to the fibers as seen on
Figure 114 and are then followed by multiple peaks as the subsequent layers fail and the punch
penetrates the top skin.

Figure 114: Typical failure at first peak of PSS test

After penetration of the top skin, the punch travels through the core by propagating a crack
outwards in a conical shape. During this, the stiffness rises both due to the core being compressed
below the punch but also due to the crack moving to a larger diameter so that more crack needs to
be created per distance of downwards movement of the punch. The cross-section of the core after
penetration are seen on Figure 115, which is from a PSS test of only the foam. A major difference is
however that in the test shown in Figure 115, the diameter of the final penetration is that of the
punch, while for the laminated specimen it is the diameter of the hole in the supporting fixture.
After punching through the core, the bottom skin is pressed out through the hole in the supporting
plate. The failure that causes the peak as the bottom skin fails cannot be seen but it is noted that
after the test, all layers have usually delaminated from each other or in pairs of one 0 and one 90
degree layer together.

Page 118 af 125


Figure 115: Cross-section of PSS test of only M80 foam

The load-displacement graphs of the tested laminated specimen are shown below, where it is seen
that the behavior are fairly consistent but that there are still differences with some specimen failing
earlier than the others with the same layup. Specimen 30 did not punch through the top skin and
instead, the whole core yielded and the load maxed out the capacity of the test machine. As the
failures of the top skin seems to be more related to bending than to through-thickness shear, it may
be that the failure loads would be higher if the test specimen were larger.

Figure 117: Load-displacement graph of PSS test, 4 layer


Figure 116: Load-displacement graph of PSS test, 4 layer skin,
skin, 25mm M130 core
25mm M80 core

Page 119 af 125


Figure 118: Load-displacement graph of PSS test, 4 layer
skin, 10 mm M80 core Figure 119: Load-displacement graph of PSS test, 8 layer skin,
25mm M80 core

Figure 120: Load-displacement graph of PSS test, 14 layer


skin, 25mm M80 core

Modelling of the penetration through the core requires modelling the crack front as it propagates.
Such a model requires input data which cannot be extracted from the tests which have been
conducted in this project and such, it has been deemed out of scope of this project and thus only the
modeling of the first peak is attempted.
Production of a model capable of accurately predicting the failure load during perimeter shear stress
testing has not been possible.
Using solid-shell elements for the core results in a distorted mesh which cannot be refined and thus,
that model is not of any use.
Using shell elements for the skins does not result in failures with clear drops in load but just very
large deformations and remeshing of the core. Remeshing of the core when the mesh is highly
distorted will come with a drop in load which can easily be misinterpreted as the first peak but this
is an artifact of the modeling.

Page 120 af 125


Using solid-shell elements for the skins have yielded similar results as when shell elements are
used, where exceeding the failure load of the skin elements does not lead to noticeable drops in load
and the simulation runs until a distorted mesh results in failure to converge.
As all simulation attempts have failed to produce a noticeable drop in load, no simulation data is
produced.
The following part of this subsection is added in V2.2, the 20th of August 2020
In Table 29, the mean values of the PSS tests are given along with the spread as a percent of how
much the datapoint furthest from the mean is. For the 14-layer skins, the specimen which didn’t
penetrate at 50 kN is omitted as this is due to the yielding of the core which makes the strength
closer to a 28-layer specimen than a 14-layer specimen after the complete collapse of the core.
4-layer skins 8-layer skins 14-layer skins 4-layer skins 4-layer skins
25mm core 25mm core 25mm core 25mm core 10 mm core
M80 foam M80 foam M80 foam M130 foam M80 foam
First peak, 7523 10443 12335 10660 8271
Newton +/- 0,6% +/- 1,1% +/- 1,1% +/- 4,1% +/- 6,5%
Second peak, 11131 20970 42175 13587 10657
Newton +/- 16,9% +/- 9,4% +/- 4,8% +/- 3,4% +/- 2,1%
Table 29: mean forces for first and second peak of the PSS tests along with % spread

Without a model that simulates the physics of the PSS test, some predictive capability can be gained
by building an empirical model by fitting some mathematical function to the test data, similarly to
how the Pacejka “magic formula” is used to model tire friction without being based on any physics.
Building such a model would require a large dataset and modelling how the different parameters
interact would require an immensely large dataset.
As there are only two different core thicknesses and two different core densities tested, these are
assumed to be linear scalings of the function for the skin thickness, as two datapoints can’t be used
to uniquely define very many functions. Linear functions are however only approximately correct at
interpolation or very low levels of extrapolation as physical phenomena are never truly linear.
The modelling of the load as a function of the skin thickness should have a value of zero at 0 layers,
it should be monotonically increasing but with a decreasing first derivative which goes to 0 at
infinity. This is assumed from the fact that adding one more layer increases the total thickness a lot
when there are few layers and increases it only a little when there are many. The necessary
asymptotic nature is explained by the fact that the compressive stress will at some point limit the
load carrying capacity. Sigmoid functions are a class of functions which could yield this capability
but at the current date (20th of August), no suitable function has been determined.
The fact that the 2nd peak of the 4-layer M80 core laminate schemes overlap means that the linear
relationship that can be formed by this data should not be considered reliable. Dependency on the
core thickness is omitted for the 2nd peak, also as a reduction in strength as the core thickness goes
to zero does not intuitively make sense.

Page 121 af 125


5.5 Optimization and recommended layup
To find the limits to be used for the optimization, the error found in section 5.3 is added to the
requirements from the rules/SES to find the constraints for the optimization calculations. As
For the beam model, the calculated stiffness must exceed the requirement by 35%, the energy
absorption must exceed half of the requirement and the maximum load must exceed the requirement
exactly.
For the 2D beam model, the calculated stiffness must exceed the requirement by 27%, the energy
absorption must exceed 55% of the requirement and the maximum load must exceed 90% of the
requirement.
The full model in ANSYS is considered the last step before physical tests which only few layups
will be subjected to. Therefore, special care should be taken to ensure that the solution is mesh
independent and time step independent and the requirements should be adjusted based on further
physical testing and development of the modelling capabilities. In this project, the recommendation
will be based on a layup in which the stiffness, energy absorption, yield load and maximum load all
exceed the requirements by 5% when simulated in this way.
For the sides of the side impact structure, the requirements as required by the rules are an equivalent
bending stiffness of 𝐸𝑒𝑞 𝐼𝑙𝑎𝑚 = 3,406 ⋅ 109 [𝑁𝑚𝑚2 ] as calculated by (3-1) and (3-2). The laminate
stiffness as simulated is corrected using equation (4-3) to be used for the equivalency calculation.
For calculation of 𝐼𝑙𝑎𝑚 , the height of the side impact structure is taken to be 250 mm. The
requirements for energy absorption, yield strength and maximum strength are found in section 4.3.1
to be 86,47 Joule, 3767,7 Newton and 8735,6 Newton respectively.
For the optimization, the fiber directions are chosen so the layers furthest from the center on each
side is laid in the fiber direction until 50% of the layers have been filled, after which the remaining
layers are placed at alternating + and – 10 degrees as required by the rules.
With this constraint on the fiber orientation and using the beam model with core properties from the
data sheets, all possible layups with Gurit Corecell M-series cores are tested with core thickness
from 2,5 to 30 mm in 2,5 increments and with the skin thicknesses being all combinations from 2 to
8 layers. A total of 2940 combinations are simulated in 16,8 seconds and the stiffnesses and
specific stiffnesses are shown in Figure 121 and Figure 122 respectively, where the red dots are
those that do not have sufficient stiffness, the blue dots are those which have sufficient stiffness and
energy but not the maximum load, the magenta dots are those who has sufficient stiffness and max
load but not sufficient energy and the green dots are those which meet all of the requirements.

Page 122 af 125


Figure 121: Stiffness of laminates by beam model for Figure 122: Specific stiffness of laminates by beam model for
optimization purposes optimization purposes

The lightest usable laminate is found to have 2 layers of SE84LV-HEC on either side of a 27,5mm
thick core of M100 foam and a total mass of 4,837 kg per m2. Among the 50 lightest usable
laminates are no layups with M60 or M200 foams, there are none with a core thickness of less than
17,5mm, and there are no layups with more than 8 layers.
Removing the laminate components that were not found among the 50 lightest usable laminates
from the 1D beam model reduces the layups to be analyzed to 450 combinations, which takes 10
minutes to calculate as a beam with a 2D core. Of these, only 88 are usable as most of the laminates
have insufficient stiffness. Of the usable lamiantes, most are with an M130 core, as opposed to
when the beam model are used for optimization where only very few of the lightest usable layups
had M130 cores. This seems to indicate that the core stiffness is a limiting factor to the optimization
and thus the M200 core are added back in, the M80 core is removed and the minimum core
thickness is reduced to 15mm to allow for layups with a thinner but stiffer core.
525 laminates are calculated as beams with a 2D core in 11 minutes but the 6 lightest layups are still
with M130 cores before M200 and M100 cores are among the usable layups. The lightest layup is 3
layers for top skin and two layers for bottom skin with a 30 mm thick M130 core and a mass of
6,551 kg per m2.
The optimum layup from the 2D beam model is calculated to have a stiffness of 2986 N/mm and is
brought on into ANSYS to be calculated at a higher level of detail. It is first analyzed linearly with a
unit displacement load to determine if the stiffness is sufficient. At a mesh with a 5mm face size
and 8 elements through the thickness of the core, the stiffness is 2976,3 N/mm and when refined to
a face size of 3mm, the stiffness rises to 3119,9 N/mm which is a 5% rise from the coarser mesh.
Refined to a face size of 2mm and 15 elements through the thickness of the core, the calculation
time is unacceptably long and the previous result is taken as having a suitable level of refinement.
As the stiffness is calculated approximately equal to that of the 2D beam simulation which is almost
30% above the rules requirement even before equation (4-3) is used to correct the stiffness,
requirement in the 2D beam optimization simulation is lowered to be only 10% higher than the
amount required by the rules. The result of the optimization search is shown in Figure 123 and
Figure 124 where it can be seen by the magenta dots that insufficient energy absorption becomes as

Page 123 af 125


much a limiting factor as the stiffness is. As noted earlier, the calculation of the energy absorption is
not very accurate due to lacking nonlinear modeling, so it is possible that the lightest rules
compliant structure is not among the green dots here.

Figure 123: Stiffness of laminates by 2D beam model for Figure 124: Specific stiffness of laminates by 2D beam model
optimization purposes for optimization purposes

The next iteration of simulation have the 4 layers as the top skin with the layup
[0°, 0°, 25𝑚𝑚 𝑀80 𝑐𝑜𝑟𝑒, 10°, −10°, 10°, 0°]. The mass is 6,3217 kg/m2 and the stiffness is
calculated by the 2D beam model to be 2678,7 N/mm and by linear analysis in ANSYS to be
2470,8 N/mm. When corrected by (4-3), the stiffness to be used by the SES is 3259,7 N/mm and the
calculated equivalent bending stiffness, 𝐸𝑒𝑞 𝐼𝑙𝑎𝑚 , are 1,9% above the rules requirement. Rather than
iterating further, the height of the side impact structure is raised from 250 mm to 258mm so that the
equivalent bending stiffness is 5% above the rules requirement. Raising the height results in a
lighter structure than if the next lightest laminate with a higher stiffness than this is used.
The calculated energy absorption is 150,2 Joule, the yield load is 11922 N and the highest measured
load is 16808 Newton, all well beyond the rules required limits. The stiffness in the nonlinear
analysis is slightly higher than in the linear analysis and is found to be 2487,5 N/mm in average
over the first millimeter. The value chosen as the yield load is the onset of plastic deformation,
which is also marked in Figure 126 and Figure 127. On Figure 125, the deformed shape and the
shear stress through the core are shown and a notable feature is that the shear strain seems to be a
little more complex than the near-constant value which the beam model calculates. The distribution
of the shear strain is slightly different than what is seen in simulations of the symmetric laminates
but this is likely caused by the stiffer top skin as the plastic strain is much less concentrated below
the applied load at the center.

Page 124 af 125


Figure 125: Simulation of optimum layup in ANSYS. Elastic shear strain through the core is plotted

Figure 126: Load-displacement graph of optimum layup,


Figure 127: Stiffness graph of optimum layup, simulated by
simulated by ANSYS full model
ANSYS full model

The total area on the monocoque that needs to be laminated is 2,7835 square meters, which means
that if all of the monocoque were to be made with this laminate, the weight would be 17,6kg plus
about 5,5 kg of steel tubing for the roll hoops and the main hoop bracing, not including attachment
brackets. If structurally unregulated areas such as the firewall behind the driver’s back and the floor
beneath the driver’s legs are subtracted due to being able to be made from lighter laminates, there
are 2,175 square meters of chassis, resulting in 13,74 kg plus the steel tubing. The whole chassis
can’t however be made from this laminate as the requirements are different for different parts of the
car, such as with the accumulator protection structure and tractive system protection structure
requiring equivalency to 3 of the tubes that the side of the SIS only requires equivalency to two of.
The requirements for the front bulkhead support structure is however lower, so the masses
mentioned here are likely to be a reasonable estimate.
The process of searching for optimum layups for the rest of the monocoque is done in a similar
fashion as shown for the side impact structure but due to time constraints in this project, it cannot be
included in this report.

Page 125 af 125


6 Conclusions
In this project, the 3 point bending of flat laminated composite plates with a foam core are modeled
to provide a basis for preliminary design choices regarding a Formula Student composite
monocoque. To validate the models, physical test have been conducted on plates with layups that
allow for the influence of three parameters to be assessed. These three parameters are the skin
thickness, the core thickness and the core stiffness which are evaluated from five layups, each with
3 test specimens. To improve the modelling, material tests have also been conducted on both the
composite and the core materials to determine material properties instead of relying on the data
sheets and guesswork. Tests to determine the flexibility of the testing machines have also been
conducted to correct for this in the experimental results and calculate the stiffness of the simulated
results as if they were experimental.
A usable Timoshenko beam finite element for laminated beams have been produced in an analytical
way that yields an exact solution of the beam equation. This element is developed in the xy-plane
and reduces to the beam element that is taught to 4th semester mechanical engineering students
when the beam is not laminated. This element is used to model 3-point bending of composite plates
with a foam core which are required in the Formula Student competition when designing a
composite monocoque. The beam element predicts the stiffness relatively well but the inability to
show compression of the core in a 3-point bending makes the element overpredict the stiffness
when this deformation is significant. To allow for more accurate stiffness prediction, the beam
element is used as the skins with the core being modeled by 2D plane stress elements. To further
improve the accuracy, the commercial software ANSYS is used to model the core 3-dimensionally
and the skins 2-dimensionally and simulate with both geometrical and material nonlinearities. These
models are used in succession and iteratively as improving accuracy comes at a cost of computation
time and as the models that are evaluated in MATLAB has the advantage of being put into for-loops
to run parametric studies of anything. The 1D beam model simulates a 3-point bending in 0,05
seconds, the 2D beam model simulates a 3-point bending in 1,5 seconds and the nonlinear 3D
model in ANSYS takes 30 minutes to compute.
The developed beam models are advantageous for brute-force optimization methods where many
laminate combinations are simulated to find the optimum layup. This is done for the side impact
structure of a Formula Student monocoque where the optimum layup is found to be
[0°, 0°, 25𝑚𝑚 𝑀130 𝑐𝑜𝑟𝑒, 10°, −10°, 10°, 0°] with a bending stiffness of 2487,5 N/mm and a mass
of 6,3217 kg per square meter. The optimum layups for the rest of the car are not evaluated due to
time constraints but an estimate of the weight of the monocoque is 13,74 kg for the regulated
composite structures of the monocoque if all of it are made as the side impact structure. Including
the mass of the roll hoops but excluding any brackets, the mass of the monocoque chassis becomes
19,24kg. Compared to the 33 kg that the steel chassis of the Viking X has and the large amount of
bodywork that may be unneeded, the weight savings are likely around 15 kg.
To comply with Formula Student rules, a kind of penetration test called the perimeter shear stress
test has been conducted on the same type of layups as the 3-point bending tests with the purpose of
modeling these as well but modeling of these has been unsuccessful.

Page 126 af 125


6.1 Discussion
Among most important source of error are incorrect calculation of the testing rig stiffness as this
leads to the simulations being compared to the wrong experimental values. The calculations of the
steel plate theoretical stiffness as a Timoshenko beam is likely to slightly underestimates the
stiffness of a wide beam. If the stiffening effect at 1,0351 of a full plate width as calculated in
section 3.5, the theoretical stiffness would be 3551 N/mm instead of 3430,8 N/mm, which would
reduce the calculated testing rig stiffness by 18,6% to 18878 N/mm. As mentioned in section 4.3.1,
the tested steel plate wasn’t a full width but two plate strips laid alongside each other, which
reduces the stiffening effect by allowing more anticlastic curvature to take place but it is still a
source of error that the testing rig is possibly more flexible than the value used in this project.
Yet another important source of error are incorrect modeling of the stress-strain behavior of the core
material as it is seen to behave nonlinearly at all times. Whether this is because yielding and
damage are initiated immediately from the start of testing, if it is because the material behaves
nonlinearly elastic or if the correction of the cross-section by the constant-volume assumption is too
incorrect is not known. If all of the nonlinear strain is plastic strain as it is assumed in this project,
the plastic elongation of the M80 test specimen should be 1,5 to 2,5mm but when holding the two
halves of each specimen together and measuring the gage length marks, it seems that the plastic
strain is less than 1mm which indicates that at least some of the strain attributed to plastic
deformations are due to unaccounted nonlinearities in the elastic deformation. When the nature of
the nonlinearity is unknown, all material nonlinearities can look like plasticity during loading so as
long as there are no unloading of the material, the use of a suitable plasticity model to model
nonlinear elastic behavior should give approximately the correct result, with the main error source
being that plasticity is calculated under a constant volume assumption while the volume change for
elastic behavior is determined by the Poisson’s ratio.
The bilinear modeling of the M80 core material has shown to not be generally applicable to all
combinations of core and skin thicknesses and especially so when both the onset of yielding, the
final load and the absorbed energy are the wanted outputs. The need is for the option to start the
plastic deformation early but still end up with high stresses and a low tangent modulus at the end,
which the multilinear model is suitable for. When the model has been applied in ANSYS, the
nonlinear simulation has failed to converge much earlier than plastic deformation should even occur
and the addition of more interpolation points in the multilinear model has not helped, nor has it to
remove any. Setting smaller timesteps and finer discretization has also been without success and
thus, the task of making it work falls upon future students who wish to build upon the work of this
project. For saving money on making physical tests, making a working multilinear model to
improve accuracy of post-yield deformation should be a priority.
Almost all of the simulations of the full model in ANSYS exhibits a suspicious behavior as the
stiffness starts by falling slowly for the first millimeter, which is consistent with the experimental
results but then a sudden change in the behavior occurs and the stiffness rises until yielding of the
core as seen in Figure 127. The behavior is not found to be caused by timestepping as it persist even
when very small timesteps of constant magnitude are used and it is not found to be caused by the
discretization as reducing the element size does not influence it. The behavior is not seen when the
core is simulated without any skins so it is likely caused by some interaction between the skin and
the core meshes. The behavior persists when the skin is offset to not intersect the core material and

Page 127 af 125


visual inspection of the deformed mesh in ANSYS does not show anything suspicious. This
behavior is bound to have an effect on the results as the stiffness is overestimated for a while but
further time cannot be allocated to identify the cause as the results fit the experimental data
somewhat well.
The sharp drops in load and stiffness as laminates fail have not been observed in the simulations,
and especially crucial it is for the simulation of the PSS tests as these failed completely to yield a
usable result. To properly simulate the failure of laminates a fracture mechanical approach is
needed but the work required to produce the needed material data and set up workable models have
been deemed too much for this project which is already very extensive in regards to modelling the
3-point bending.
The optimization in this project has been focused on finding the lowest mass but since the price of
the materials are known to purchase them for the tests, an optimization regarding price can just as
easily be done. The price optimized laminate will likely be a laminate which has the least amount of
carbonfiber, which is also likely to be among the lightest as carbonfiber is significantly heavier than
the foams used for the core material.

Page 128 af 125


7 Biblography
[1] “Formula Student rules 2020”, version 1.0, available from
https://www.formulastudent.de/fileadmin/user_upload/all/2020/rules/FS-Rules_2020_V1.0.pdf
[accessed 03/08-2020]
[2] Mallick, P. (2008) ”Fiber-Reinforced Composites: materials, manufacturing and design” 3rd
edition, ISBN 0-8493-4205-8
[3] Hashin, Z., (1980) “Failure Criteria for Unidirectional Fiber Composites”, Journal of Applied
Mechanics, Volume 47, pages 329-334
[4] Przemieniecki, J. (1968) “Theory of matrix structural analysis”, ISBN 0-486-64948-2
[5] Reddy, J. (2004) “Mechanics of laminated composite plates and shells: theory and analysis”, 2nd
edition, ISBN 0-8493-1592-1
[6] Timoshenko, S., Goodier, J. (1970) “Theory of elasticity”, 3rd edition, ISBN 0-07-064720-8
[7] Budynas, R., Sadegt, A., (2020) “Roark’s Formulas for Stress and Strain”, 9th edition, ISBN
978-1-260-45375-1
[8] Ashwell, D., (1950). “The anticlastic curvature of rectangular beams and plates”. Journal of the
Royal Aeronautical Society. Volume 54. Pages 708-715
[9] Eisenberger, M., (1994) “Deriviation of Shape Functions for an Exact 4-D.O.F. Timoshenko
Beam Element”, communications in numerical methods in engineering, volume 10, pages 673-681
[10] Cook, R., Malkus, D., Plesha M., Witt, R. (2002) “Concepts and applications of finite element
analysis”, 4th edition, ISBN 978-0-471-3565-9
[11] Madabhusi-Raman, P., Davalos, J., (1996) ”Static shear correction factor for laminated
rectangular beams”, Composites Part B, volume 27, pages 285-293
[12] Whitney, J., (1987) “Structural Analysis of Laminated Anisotropic Plates”, 1987, ISBN
9780877625186
[13] Shimpi, R., Shetty, R., Guha, A., (2017) ”A simple single variable shear deformation theory
for a rectangular beam” , Journal of mechanical engineering science, volume 23, pages 4576-4591
[14] Shimpi, R., Patel, H. (2006) “A two variable refined plate theory for orthotropic plate
analysis”, International Journal of Solids and Structures, volume 43, pages 6783-6799
[15] Wilson, E., Ibrahimbegovic, A., (1990) “Use of incompatible displacement modes for the
calculation of element stiffnesses or stresses”, finite elements in analysis and design, volume 7,
pages 229-241
[16] Wang, C., Lim, G., Reddy, J., Lee, K., (2001) “Relationships between bending solutions of
Reissner and Mindlin plate theories”. Engineering Structures. Volume 23. Pages 838-849
[17] ANSYS 2019 R2 documentation, “Element reference”.

Page 129 af 125


[18] Tahani, M. (2007) “Analysis of laminated composite beams using layerwise displacement
theories”. Composite Structures. Volume 79. Pages 535-547
[19] Bathe, K. (2010) “RES.2-002 Finite element procedures for solids and structures – Nonlinear
analysis”, Massachusetts Institute of Technology: MIT OpenCourseWare.
[20] ANSYS 2019 R2 documentation, “Theory reference”.
[21] Bathe, K. (2014) “Finite Element Procedures” 2nd ed. ISBN 978-0-9790049-5-7
[22] Zhu, F., Bai, P., Zhang, J., Lei, D., He, X. (2015) “Measurement of true stress-strain curves
and evolution of plastic zone of low carbon steel under uniaxial tension using digital image
correlation”. Optics and Lasers in Engineering. Volume 65. Pages 81-88
[23] Xie, X., Li, J., Sia, B., Bai, T., Siebert, T., Yang, L. (2017) “An experimental validation of
volume conservation for aluminum alloy sheet metal using digital image correlation method”.
Journal of Strain Analysis. Volume 52. Pages 24-29
[24] Wang, Y., Jiang, J., Wanintrudal, C., Du, C., Zhou, D., Smith, L., Yang, L., (2010) ”Whole
field sheet-metal tensile test using digital image correlation”. Experimental Techniques. Volume 35.
Pages 54-59
[25] Chandrupatala, T., Belegundu, A. (2012) “Introduction to Finite Elements in Engineering”, 4th
edition, ISBN 0-273-76368-7
[26] Kreyszig, E. (2011) ”Advanced engineering mathematics”, 10th edition, ISBN 978-0-470-
64613-7

[27] Welch, B., (1947) “The Generalization of `Student's' Problem when Several Different
Population Variances are Involved”, Biometrika, Volume 34, pages 28-35
[28] Grower, M., Shaw, R., Broughton, W. (2008). ”NPL Report Mat 24”. National Physical
Laboratory. ISSN 1754-2979

Page 130 af 125

View publication stats

You might also like