You are on page 1of 132

ISSN 1831-9424 (PDF)

ISSN 1018-5593 (Printed)

Development of Improved
Shear Connection Rules
in Composite Beams
(DISCCO)

Research and
Innovation EUR 28458 EN
EUROPEAN COMMISSION
Directorate-General for Research and Innovation
Directorate D — Industrial Technologies
Unit D.4 — Coal and Steel

E-mail: rtd-steel-coal@ec.europa.eu
RTD-PUBLICATIONS@ec.europa.eu

Contact: RFCS Publications

European Commission
B-1049 Brussels
European Commission

Research Fund for Coal and Steel


Development of Improved Shear
Connection Rules in Composite Beams
(DISCCO)

R. M. Lawson and E. Aggelopoulos


SCI
Silwood Park, Ascot, Berkshire, SL5 7QN UNITED KINGDOM

R. Obiala and F. Hanus


ArcelorMittal Belval and Differdange S.A.
L-4008 Esch-Belval and L-4503 Differdange, LUXEMBOURG

C. Odenbreit and S. Nellinger


University of Luxembourg
Campus Kirchberg, 6, rue Coudenhove-Kalergi, L-1359 Luxembourg-Kirchberg, LUXEMBOURG

U. Kuhlmann and F. Eggert


University of Stuttgart
Pfaffenwaldring 7, 70569 Stuttgart, GERMANY

D. Lam, X. Dai and T. Sheehan


University of Bradford
BD7 1DP Bradford, UNITED KINGDOM

Grant Agreement RFSR-CT-2012-00030


1 July 2012 to 30 June 2015

Final report

Directorate-General for Research and Innovation

2017 EUR 28458 EN


LEGAL NOTICE
Neither the European Commission nor any person acting on behalf of the Commission is
responsible for the use which might be made of the following information.
The views expressed in this publication are the sole responsibility of the authors and do not
necessarily reflect the views of the European Commission.

Europe Direct is a service to help you find answers


to your questions about the European Union

Freephone number (*):


00 800 6 7 8 9 10 11
(*) Certain mobile telephone operators do not allow access to 00 800 numbers or these calls may be billed.

More information on the European Union is available on the Internet (http://europa.eu).

Cataloguing data can be found at the end of this publication.

Luxembourg: Publications Office of the European Union, 2017

Print ISBN 978-92-79-65674-3 ISSN 1018-5593 doi:10.2777/1186 KI-NA-28-458-EN-C


PDF ISBN 978-92-79-65673-6 ISSN 1831-9424 doi:10.2777/923858 KI-NA-28-458-EN-N

© European Union, 2017


Reproduction is authorised provided the source is acknowledged.
Contents
Page No

1 Final Summary 5
1.1 Project objectives 5
2 Scientific and technical description of results – WP1 9
2.1 WP1: Realistic tests to evaluate shear connector resistances 9
2.2 Task 1.1: Evaluation of existing push-out tests for shear connectors in
composite slabs 9
2.3 Tasks 1.2-1.3: Push-out tests and recommended standard test regime 10
2.4 Influencing parameters 23
2.5 Comparison with characteristic resistances to EN 1994-1-1 31
2.6 Recommended standard test regime 32
2.7 New mechanical model for shear connector resistance based on this research 33
2.8 Comparison of initially planned activities and work accomplished 35
3 WP2: Composite beam tests to investigate the influence of low degrees of shear
connection 37
3.1 Objectives of WP2 37
3.2 Description and impact of the research results 38
3.3 Development of end slip 45
3.4 Comparison of beam tests with plastic bending resistance for partial shear
connection 46
3.5 Concluding remarks for WP2 48
4 WP3: Tests on long span asymmetric composite beams 49
4.1 Objectives of WP3 49
4.2 Comparison of initially planned activities and work accomplished 49
4.3 Description and impact of the research results 49
4.4 Comparison of deflections of the tests 57
4.5 Concluding remarks for WP3 59
5 WP4: Composite floor plate test to investigate interactive effects 61
5.1 Objectives of WP4 61
5.2 Description and impact of the research results 61
5.3 Floor plate results 65
5.4 Results of robustness test 70
5.5 Modelling of floor plate test 71
6 WP5: Numerical models for composite beam analysis using shear connector properties 75
6.1 Objectives of WP5 75
6.2 Description and impact of the research results 75
6.3 WP 5.2 Modelling of composite beams 82
6.4 WP5.3 Modelling of cellular beams with low degrees of shear connection 84
6.5 Concluding remarks for WP5.3 91
6.6 WP5.5 Simplified finite element models 91
6.7 Concluding remarks for WP 5 93
7 WP6: Improved shear connection rules for composite beams 95
7.1 Objectives of WP6 95
7.2 Comparison of initially planned activities and work accomplished 95
7.3 Description and impact of the research results- WP6.1 96
7.4 WP6.2 Rules for partial utilisation 102
7.5 Beams loaded by point loads 108
7.6 Beams loaded with heavy loads 110
8 WP7: Project management and coordination 113
8.1 Objectives of WP7 113
9 Exploitation Plan 115
References 117
10 LIST OF FIGURES 121
11 LIST OF TABLES 125
12 LIST OF ACRONYMS AND ABBREVIATIONS 127
3
4
1 Final Summary

1.1 Project objectives


Composite construction leads to efficiencies in use of materials and to improved structural
performance. For composite beams, rules for partial shear connection between the concrete slab
and the steel section is included in the current EN 1994-1-1: Eurocode 4. However, the existing
rules for the minimum degree of shear connection, in some cases, make the design of composite
beams impossible to satisfy, particularly for the use of modern deck profiles with widely spaced ribs
and hence fewer shear connectors.

The key objective of the DISCCo project was to develop new rules for the partial shear connection
of composite beams, which will lead to improvements in EN 1994-1-1: Eurocode 4. A series of
tests and analyses were undertaken with the following objectives:

 To develop a new standard push-out test that will be more appropriate for composite slabs
rather than solid slabs. This is based on a modified push-test with transverse compression
applied as a proportion of the in-plane shear force.
 To obtain more information about the load-slip characteristics of shear connectors in the
troughs of decks and investigate a number of parameters affecting their resistance.
 To investigate the performance of composite beams with low degrees of shear connection. Short
span beam tests were performed in which approximately one—third of the span was subject to
constant shear, and which conformed to the same deck profile, shear connector arrangement
and reinforcement as the push-tests.
 To investigate the behaviour of long span composite and cellular composite beams with low
degrees of shear connection and a high asymmetry in the cross-sectional shape to extend the
performance data into the full range of practical application.
 To investigate the membrane effects in the slab in a grillage of composite beams that reduce
the tendency for longitudinal splitting along the line of shear connectors in a beam, where steel
decking is orientated parallel to the beam and is discontinuous across it. This also extended to
the performance of edge beams with and without additional U-bar reinforcement.
 To develop numerical models that will be able to analyse composite beams with greater
accuracy and obtain more realistic predictions. Finite element modelling was performed using
ABAQUS and Ansys on solid web beams of symmetric and asymmetric section and cellular
beams with low degrees of shear connection.
 Development of appropriate design rules based on both the test and FEA results.
This research is comprehensive in both its testing and analysis and has led to practical guidance
suitable for incorporation in the revision to EN 1994-1-1.

WP1: Realistic tests to evaluate shear connector resistances


A total of 70 push-tests was performed at the University of Luxembourg and Stuttgart to an agreed
test regime. The Luxembourg tests used IPE 360 beams and an 80 mm deep trapezoidal deck
profile with ribs at 300 mm spacing, which is typical of modern profiles for longer spans. The
Stuttgart tests used IPE 300 beams and a 56 mm deep trapezoidal profile with ribs at 207 mm
spacing, and also a small number of tests on re-entrant decking. The transverse loading was
varied between 0 and 15% of the applied shear loading and for the deeper deck profile was shown
to increase the deformation capacity but had a modest effect on the shear resistance. Transverse
applied moment had little additional effect. A new method of performing a push-test is proposed.

For the deeper deck profiles, the mode of failure was dominated by a concrete failure ‘cone’ over
the shear connectors; whereas for the shallower profiles, an ‘S-shaped’ bending failure of the shear
connectors was observed. This leads to the development of a new model for shear connector
resistance.

WP2: Composite beam tests to investigate the influence of low degrees of shear
connection
A total of 10 composite beams was tested, of which 8 were 5m span using the 56 mm deep deck
profile and IPE300 beams, and 2 were 6m span using the 80 mm deep deck profile and IPE360
beams. The degree of shear connection was less than 40% in all but one case. The same shear
connector arrangements as in WP1 were used in order that the characteristic shear connector
resistance could be input into the plastic stress block analysis ‘model’.

5
Nine of the beam tests gave an average ratio of failure load: predicted failure load of 0.98, when
using a shear connector characteristic resistance of 0.9 x test resistance. This is considered to be a
very good comparison and justification of the Eurocode 4 plasticity model, despite the low degrees
of shear connection. One test with pairs of shear connectors failed at about 15% lower load than
predicted because of poor weld quality. This highlights the risk of placing pairs of shear connectors
close to the edge of the flanges.

This test series led to the development of new rules for the influence of the transverse deck profile
on shear connector resistance.

WP3: Tests on long span asymmetric composite beams


The purpose of these tests on long span beams was to investigate the behaviour of modern
composite construction systems designed for low degrees of shear connection. The following tests
were performed:

 Test on 15.3m span cellular beam with regular circular openings and with asymmetry in flange
areas of 2.4:1. The beam had 36% degree of shear connection for single shear connectors per
deck rib, which was well below the Eurocode 4 requirement of 85%.
 Test on cellular beam with an elongated opening subject to shear which relied on the degree of
shear connection developed over the opening.
 Test on 11.2m span fabricated beam with asymmetry in flange area of 1.5:1. The fabricated
beam test used the same shear connector arrangement and deck profile as in the cellular beam
test, and had 33% degree of shear connection for single shear connectors per deck rib.
In all of the tests, the calculated plastic bending resistance using the shear connector resistance
obtained from push-tests was within 2% of the failure moment, despite the low degree of shear
connection that was less than half of that required by Eurocode 4 for these long span beams. The
end slips at failure ranged from 15 to 20 mm at failure which far exceeded the limiting end slip of
6m in Eurocode 4, and so justify a reduction in the degree of shear connection rules.

The deflections of the long span beams at serviceability loads agreed well with the prediction using
the elastic stiffness for partial shear connection developed in WP6.1. A simple formula for the
additional deflection due to the openings in cellular beams was developed.

WP4: Composite floor plate test to investigate interactive effects


A floor plate of 10m x 4m plan area was constructed using IPE270 and IPE300 beams supported by
6 HEA 200 columns. Four tests were performed:

 Loading of internal secondary beam to cause failure of the supporting primary beams
 Loading of edge beams with U-bar reinforcement around the shear connectors.
 Loading of edge beams without additional reinforcement.
 Pseudo-robustness test by removing support of the internal column, and applying load to the
floor plate.
The test confirmed that primary beams reached their plastic bending resistance despite the use of
discontinuous decking and transverse reinforcement of the minimum percentage given in Eurocode
4. The tests also led to a proposal for a reduction factor due to shear connectors at edge beams
without U-bars.

WP5: Numerical models for composite beam analysis using shear connector
properties
Finite element models of the test beams in WP2 and 3 were created using ABAQUS and ANSYS
using shear connector load-slip relationships based on the results of WP1. It was agreed that the
elastic behaviour for single shear connectors may be based on a stiffness of 70 kN/mm, and for
pairs of shear connectors on a stiffness of 100 kN/mm. A declining part of the load-slip curve was
also introduced after a maximum slip of 6 mm.

FE models were also set up for IPE300 to IPE750 beams of representative spans to investigate the
minimum degree of shear connection to achieve 6 mm end slip for both propped and un-propped
beams. It was found that the minimum degree of shear connection may be reduced by 15 to 30%
for unpropped beams relative to propped beams.

6
FE models were also set up for cellular beams of 9 to 20 m span, which were compared to the
same beams with solid webs. It was found that the minimum degree of shear connection may be
reduced for cellular beams, but this effect was less for highly asymmetric profiles than for
symmetric profiles.

The numerical models also extended to an investigation of local shear connection behaviour. This
was found to be more complex because of the use of Explicit FE models with concrete splitting.

WP6: Improved shear connection rules for composite beams


Improved shear connection rules were developed based on the results of the extensive test
programme and FE model as follows:

 New reduction factor formula for the effect of the deck profile shape and shear connector
height/diameter on the shear connector resistance.
 Rules for partial shear connection of un-propped beams and asymmetric beams.
 Rules for partial shear connection of cellular beams, expressed as a modification factor to be
equivalent to solid web beams.
 Rules for additional deflection due to partial shear connection, and due to the regular openings
in cellular beams.

WP7: Project management and coordination


The Consortium of DISCCo partners worked effectively together and the extensive test series was
completed to time and to a consistent quality of output. The partners also developed suitable FE
models which were exchanged to facilitate rapid progress in WP5 in particular.

7
8
2 Scientific and technical description of results – WP1

2.1 WP1: Realistic tests to evaluate shear connector resistances

2.1.1 Objectives of WP1


The objective of this work package was to develop a standard push-out test aiming to replace the
existing test in EN 1994-1-1, which was developed for solid slabs and not for composite slabs.
Having established a representative push-out test, a series of parametric tests was carried out in
order to investigate the influence of a range of parameters such as the deck profile shape and
depth, reinforcement pattern, concrete strength, etc. on shear connector resistance and
deformation capacity. It was then possible to categorise shear connectors in composite slabs
according to their load-slip characteristics.

The following tasks were completed according to the original proposal:

Task 1.1 Evaluation of existing push-out tests for shear connectors in composite slabs

Task 1.2 Development of standard push-out test for shear connectors in composite slabs

Task 1.3 Discussion on push-out test results and investigation of influencing parameters on
shear connector resistance and ductility

2.2 Task 1.1: Evaluation of existing push-out tests for shear connectors in
composite slabs
Information on existing push-out tests was collected by UStutt and SCI. UStutt summarised the
data from European studies, some of which were evaluated in the thesis of Konrad. SCI gathered
data from push-out tests carried out mostly in Australia and the USA. The results from these tests
were then compared against predictions using the EN 1994-1-1 equations for the shear connector
resistances (see Figure 2.1).

Figure 2.1 Shear connector resistances from tests compared to characteristic resistance to EN
1994-1-1

F=favourable, U=unfavourable position, 2=pairs per rib, parallel= decking parallel to beam)

Previous push tests had shown that a transverse load of 12% of the applied shear load would lead
to improved performance where dominated by concrete failure and is justified by consideration of
the loading applied through the slab for secondary beams. It was also found that shear connectors
exhibited high ductility and recorded slips were found in excess of the 6 mm limit in Eurocode 4
(see Figure 2.2). A short investigation into the parameters affecting the SCI test results and a
statistical analysis to EN 1990 were also carried out.

9
SCI push-out test data vs EN1994:1-1
140 60-1C
Ptest (kN/stud) 120 60-2C
100
80-1C
80
60 80-2C
40 60-1F-TS
20
60-1F-TD
0
0 20 40 60 80 100 120 140
PRk (kN/stud)

Figure 2.2 Shear connector resistances obtained from tests by SCI compared with characteristic
resistance to EN 1994-1-1 (C=central position, F=favourable position, TS=top of
slab, TD= top of deck)

Characteristic SC resistance vs slip


100
PRk (kN/stud)

80 Smith &
60 Couchman,2010
Smith,2009
40
Rackham,2008
20
0 EN 1994-1-1 limit
0 20 40 δ (mm)

Figure 2.3 Slips obtained from tests by the SCI compared with EN 1994-1-1 requirement.
Transverse loading equivalent to 12% of the test load

A comprehensive test series was performed in this WP to investigate the role of parameters such
as the magnitude of transverse loading, the amount and position of reinforcement, the number and
size of shear connector, the concrete strength, the conditions of lateral restraint, etc. Finally, a
standard push-out test regime was developed.

2.3 Tasks 1.2-1.3: Push-out tests and recommended standard test regime
A total of 28 push-out tests was performed at the University of Luxembourg and a further 36 at the
University of Stuttgart by varying the following parameters:

 type of slab: solid or composite slabs;


 type of profiled decking: 56mm re-entrant, and 58mm and 80mm deep trapezoidal profiles;
 welding process for the shear connectors: through deck welded or with pre-punched holes;
 number of shear connectors per trough: one or two at central position;
 diameter of the shear connector: 19 mm or 22 mm;
 influence of the position of reinforcement: single layer on top of the slab or two layers;
 concrete strength: C30/37 for most of the tests, C20/25 and C40/50 for some comparison
tests;
 influence of concentric and eccentric transverse loading (with and without additional transverse
load and negative bending moments) – only ULux tests;
 boundary conditions: tension ties with different types of fixation
The test results are presented and evaluated for their characteristic resistance and slip capacity
according to the simplified procedure of EN 1994-1-1 Annex B2 and the influence of all the above-
mentioned parameters is discussed.

10
Based on the observations in the push-out tests, recommendations for a more realistic standard
test regime are given with regard to the specimen geometry and the application of transverse
loading.

2.3.1 Application of transverse loading in push-out tests


For ‘ductile’ shear connectors (slip capacity of at least 6mm), EN 1994-1-1 permits use of a plastic
distribution of shear forces in which all of the studs are assumed to be loaded to their design
resistance, PRd. In a typical secondary beam with transverse decking, loaded with a distributed
load, a group of studs per deck rib is exposed to a transverse load Pr,Ed. The resulting forces acting
on the shear connectors are shown in Figure 2.4.

(a) Loading applied normal to the slab


where the decking is transverse to
the beam span

(b) Shear and normal forces acting on


the shear connectors

Figure 2.4 Forces acting on shear connectors in a composite beam

The degree of transverse loading may be defined as the ratio of the transverse load acting per deck
rib to the resistance of the studs (where nr is the number of studs in a deck rib).

1 𝑃r,Ed (2.1)
𝜈d = ∙
𝑛r 𝑃Rd
Where Pr,Ed is the applied normal load from the slab expressed per deck rib

and PRd is the design resistance of a shear connector with transverse decking

In a push test, the slab is attached on both sides of the steel profile, and so the applied force in the
test is twice the longitudinal shear resistance of each side of the slab. Therefore, the transverse
shear force as a proportion of the shear force per side is twice the ratio of transverse load to
applied load in the test.

A parametric study was conducted to determine the sensible ratio of transverse to longitudinal
shear forces assuming that the shear connectors reach their design resistance. The span of the slab
was varied between 3 m and 6 m for 56 mm and 80 mm deep deck profiles (although the practical
range is 3 to 4.5 m). The span of the composite beam was varied between 8 m and 16 m. For
each span configuration, the lightest possible steel profile was chosen which satisfies the bending
resistance and minimum degree of shear connection. The steel grade was chosen as S235 and
S355. The stud resistance was calculated according to EN 1994-1-1 Clause 6.6.2.

The results show that the ratio of slab vertical force versus the longitudinal shear force lies
between approximately 4% and 12%. Thus, the degree of transverse loading that may be applied
to the push test is between about 8% and 25% of the test load. To prevent an over-prediction of
the test load, it was decided to conduct the push-out tests at 5% and 10% transverse load. More
information on this parametric study can be found in deliverable D1.2 available in CIRCABC.

2.3.2 Push-out test setup


For the application of transverse loading, the test rig shown in Figure 2.5 was used. The test rig
was developed to load both slabs of the push-out test with only one hydraulic jack. The assembly
was designed to be modular to match the different geometries of specimens with 56mm and 80mm
deep deck profiles and 130 mm and 150 mm deep slabs and to allow the application of either
concentric or eccentric transverse loads.

11
Figure 2.5 Test set-up for the application of transverse loading

To prevent the horizontal displacement of the concrete slabs and restraint by friction at the
supports, a layer of Teflon was placed at the supports. A 1mm thick flat steel sheet was placed on
top of the Teflon. The specimens were then placed in a mortar bed on top of the steel sheets.
Specimens tested at UStutt were accurate enough to place them directly on the steel plate of the
test machine. The applied degrees of transverse loading were chosen to have values of 5% to 16%
of the applied shear force. This range was identified as the realistic range of transverse loading in a
preliminary parametric study as mentioned above.

The two shapes of trapezoidal deck profiles are shown in Figure 2.6. The push-out test in Task 1.3
(see Table 2.1) and some tests in Task 1.2 (see Table 2.2) were conducted without transverse
loading. The test set-up used at the University of Luxembourg did not use tension ties to
horizontally restrain the specimen (see Figure 2.7). The University of Stuttgart applied tension ties
as shown in Figure 2.8. According to the observations during testing, these tension ties were not
activated.

Special cases to investigate the influence of tension ties were considered in test series 1-08 and
3-01. The tension ties were point fixed close to the corners of the slabs using different types of
fixations as shown in Figure 2.7 and Figure 2.8. Series 1-01 to 1-02-2 were performed on solid
slab specimen, while series 3-03 to 3-10 used 56mm deep re-entrant decking (see Table 2.1 and
Table 2.2).

(a) Cofraplus 60 decking profile

(b) Comflor 80 decking

Figure 2.6 Details of the trapezoidal deck profiles

12
Figure 2.7 Details of the trapezoidal deck profiles

Figure 2.8 Test setup used at the University of Stuttgart

Figure 2.9 Typical dimensions of specimens with Cofraplus 60 decking

13
Figure 2.10 Typical dimensions of specimens with ComFlor 80 decking

Figure 2.11 Typical dimensions of specimens with Cofrastra 56 decking

In addition, a series of push-out tests with the decking placed parallel to the beam was conducted
to accompany the floor plate test of WP4. Series KL used Cofraplus 60 decking, which was not
continuous above the beam. Two shear connectors with 19 mm diameter and 100 mm height were
welded directly to the beam on each side at a spacing of 200 mm. A Q188 wire fabric (188 mm2/m
in both directions) was placed 30 mm above the decking. To simulate the behaviour of primary
beams, transverse loading was not applied in these tests.

14
Figure 2.12 Dimensions of specimens with steel decking parallel to the beam

2.3.3 Test procedure


All tests were loaded according to the procedure given in EN 1994-1-1. A preloading procedure of
25 load cycles between 5% and 40% of the expected test load was applied to all tests.

For the first test of each group (specimens 1-04-1 and 1-08-1), the specimen was loaded in steps
of 10 kN until 40% of the expected test load was reached. After each load cycle the specimen was
unloaded. Tests were carried out under displacement control and the displacement was applied in
increments of 1 mm at rate of 0.1 mm/s until a slip of about 7mm was reached. Some specimens
were unloaded at this point to gather additional information on the elastic stiffness.

When the specimens were reloaded, the displacement was applied in 2 mm increments with
0.02 mm/s until a slip of 15mm was reached. Then the increments were increased to 5 mm with
0.05 mm/s. After each load step, the displacement was kept constant for at least 5 minutes to
determine the influence of relaxation. The tests were stopped when either stud failure was
observed or the test load had fallen to less than 70% of the first load peak.

2.3.4 Typical load-slip behaviour and observed failure modes


A first series of tests was conducted with solid slabs. These tests were used as a basis for all
further tests using steel decking to investigate and determine the influence of any parameter
compared to the solid concrete slab. All performed tests had a relatively high measured concrete
strength with cylinder strengths of 44-46 N/mm² (see Figure 2.13). Therefore, the first range of
the load-slip curve was very ductile, but then a sudden steel failure (shearing of the shank of the
headed stud) was observed.
Load-slip curve 1-01 P1
1800

1600

1400 1161,0 kN 1263,5 kN

1200
Load [kN]

1000

800

600

400
10,40 mm

200

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
Slip [mm]

Figure 2.13 Typical load-slip curve for solid slab (specimen 1-01-1)

15
Table 2.1 Data for push-out tests in Task 1.2
Series i Decking Studs Concrete Reinforcement Loading
hp bm t d hsc nr fu fcm Ecm Bottom Top TL e
[mm] [N/mm²] [kN] [mm]
1 - - - 19.11 123.4 - 467 44.2 34350+ 10/150 10/150 - -
1-01 2 - - - 19.05 122.5 - 467 44.3 34390+ 10/150 10/150 - -
3 - - - 19.11 123.2 - 467 44.8 34504+ 10/150 10/150 - -
1 - - - 22.19 123.5 - 514 45.6 34673+ 10/150 10/150 - -
1-02-1 2 - - - 22.29 124.1 - 514 45.5 34662+ 10/150 10/150 - -
3 - - - 22.28 123.9 - 514 44.2 34368+ 10/150 10/150 - -
1 - - - 22.28 123.5 - 514 46.2 34824+ 10/150 10/150 - -
1-02-2 2 - - - 22.23 123.5 - 514 45.8 34718+ 10/150 10/150 - -
3 - - - 22.23 123.3 - 514 45.4 34628+ 10/150 10/150 - -
1 58 81.5 0.88 22.14 123.3 1 514 41.0 33605+ Q188A Q335A - -
1-03 2 58 81.5 0.88 22.10 122.0 1 514 42.5 33968+ Q188A Q335A - -
3 58 81.5 0.88 22.22 122.7 1 514 42.9 34061+ Q188A Q335A - -
1 58 81.5 0.88 22.16 124.3 1 551 30.1 20900 Q188A Q335A 4.1% -
1-04 2 58 81.5 0.88 22.24 124.0 1 551 30.9 21500 Q188A Q335A 12.5 -
3 58 81.5 0.88 22.20 123.9 1 551 30.9 21500 Q188A Q335A 12.5 -
1 58 81.5 0.88 22.16 124.0 1 551 30.7 22100 Q188A Q335A 8.2% -
1-05 2 58 81.5 0.88 22.17 123.8 1 551 30.7 22100 Q188A Q335A 25.0 -
3 58 81.5 0.88 22.20 123.9 1 551 32.6 22800 Q188A Q335A 25.0 -
1 58 81.5 0.88 22.18 123.5 1 551 29.9 21200 Q188A Q335A 4.1% 380
1-06 2 58 81.5 0.88 22.16 124.1 1 551 31.1 21400 Q188A Q335A 12.5 380
3 58 81.5 0.88 22.12 123.7 1 551 44.0 25900 Q188A Q335A 12.5 380
1 58 81.5 0.88 22.12 123.2 1 551 45.1 26000 Q188A Q335A 8.2% 380
1-07 2 58 81.5 0.88 22.14 124.6 1 551 40.9 27500 Q188A Q335A 25.0 380
3 58 81.5 0.88 22.14 124.2 1 511 42.6 28000 Q188A Q335A 25.0 380
1 80 137.5 0.90 19.06 117.4 2‡ 551 42.2 26500 Q188A - Weak 250
1-08 2 80 137.5 0.90 19.07 117.5 2‡ 551 42.2 26500 Q188A - Weak 250
3 80 137.5 0.90 19.07 117.9 2‡ 551 42.4 26400 Q188A - Strong 250
1 80 137.5 0.90 19.09 118.8 2‡ 551 42.6 28000 Q188A - 8.8 -
1-09 2 80 137.5 0.90 19.12 117.8 2‡ 551 42.6 28000 Q188A - 17.5 -
3 80 137.5 0.90 19.09 118.8 2‡ 551 42.6 28000 Q188A - 17.5 -
1 80 137.5 0.90 19.11 118.6 2‡ 551 42.6 28000 Q188A - 17.5 -
1-10 2 80 137.5 0.90 19.09 118.1 2‡ 551 42.6 28000 Q188A - 13.2 -
3 80 137.5 0.90 19.10 118.2 2‡ 551 42.6 28000 Q188A - - -
1 80 137.5 0.90 19.08 119.4 2‡ 551 42.6 28000 Q188A - 17.5 380
1-11 2 80 137.5 0.90 19.08 118.7 2‡ 551 42.6 28000 Q188A - 17.5 380
3 80 137.5 0.90 19.08 118.6 2‡ 551 42.6 28000 Q188A -- 17.5 380
1 80 137.5 0.90 19.09 121.3 1‡ 551 44.1 25600 Q188A - 0 -
NR1 2 80 137.5 0.90 19.09 121.2 1‡ 551 45.7 25600 Q188A - 8.8 -
3 80 137.5 0.90 19.10 121.0 1‡ 551 44.7 25600 Q188A - 17.5 -
‡ +
Through deck welded Ecm=22 (fcm/10)0.3

16
Table 2.2 Data for push-out tests in Task 1.3
Series i Decking Studs Concrete Reinforcement Loading
hp bm t d hsc nr fu fcm Ecm Bottom Top TL e
[mm] [N/mm²] [kN] [mm]
1 58 81.5 0.88 19.1 121.7 1‡ 467 42.4 33932+ - Q335A - -
2-01 2 58 81.5 0.88 19.12 121.7 1‡ 467 42.6 33991+ - Q335A - -
3 58 81.5 0.88 19.23 122.0 1‡ 467 41.8 33784+ - Q335A - -
2-02 1 58 81.5 0.88 19.00 122.5 1‡ 467 40.7 33508+ Q188A Q335A - -
2-03 1 58 81.5 0.88 19.00 121.7 1 467 39.7 33257+ - Q335A - -
2-04 1 58 81.5 0.88 19.00 120.6 2‡ 467 40.3 33423+ - Q335A - -
1 58 81.5 0.88 19.00 121.0 2‡ 467 40.2 33386+ Q188A Q335A - -
2-05 2 58 81.5 0.88 19.00 121.7 2‡ 467 38.6 33002+ Q188A Q335A - -
3 58 81.5 0.88 19.00 121.7 2‡ 467 39.2 33145+ Q188A Q335A - -
1 58 81.5 0.88 22.20 122.2 1 514 40.1 33374+ - Q335A - -
2-06 2 58 81.5 0.88 22.22 122.2 1 514 39.6 33232+ - Q335A - -
3 58 81.5 0.88 22.25 122.2 1 514 40.6 33490+ - Q335A - -
1 58 81.5 0.88 22.17 121.7 1 514 46.4 34868+ Q188A Q335A - -
2-07 2 58 81.5 0.88 22.32 122.2 1 514 46.3 34851+ Q188A Q335A - -
3 58 81.5 0.88 22.33 122.2 1 514 46.0 34774+ Q188A Q335A - -
2-08 1 58 81.5 0.88 19.09 122.0 1 467 45.8 34735+ Q188A Q335A - -
1 80 137.5 0.90 19.07 118.4 2‡ 551 46.0 26900 Q188A Q335A High 250
3-01 2 80 137.5 0.90 19.06 117.4 2‡ 551 46.0 26800 Q188A Q335A High 250
3 80 137.5 0.90 19.07 118.3 2‡ 551 40.4 26800 Q188A Q335A - -
3-02 1 80 137.5 0.90 19.08 123.4 2 551 42.6 28000 Q188A - - -
3-03 1 56 110† 1.00 19.04 122.0 1‡ 467 39.9 33312+ - Q335A - -
3-04 1 56 110† 1.00 19.12 122.2 1 467 41.0 33580+ - Q335A - -
3-05 1 56 110† 1.00 19.11 121.9 2‡ 467 39.7 33282+ - Q335A - -
3-06 1 56 110† 1.00 19.12 121.7 2‡ 467 39.5 33207+ Q188A Q335A - -
3-07 1 56 110† 1.00 19.10 122.3 2 467 40.7 33508+ Q188A Q335A - -
3-08 1 56 110† 1.00 19.14 122.00 1 467 41.3 33671+ - Q335A - -
3-09 1 56 110† 1.00 22.20 123.25 1 467 40.7 33514+ - Q335A - -
3-10 1 56 110† 1.00 22.21 122.25 1 467 40.5 33471+ Q188A Q335A - -
‡ +
Through deck welded † Re-entrant deck, top width of rib Ecm=22 (fcm/10)0.3
TL = Transverse loading
e = eccentricity of transverse load

17
Table 2.3 Parameters of tests with the decking placed parallel to the beam
Series No. Decking Studs Concrete Reinforcement
hp bm t d hsc fu fcm Ecm Top Bottom
[mm] [N/mm²]
KL 1 58 81.5 0.88 19 97 550 49.12 35465 - Q188A
2 58 81.5 0.88 19 97 550 49.68 35586 - Q188A
3 58 81.5 0.88 19 97 550 47.82 35181 - Q188A
Ecm=22 (fcm/10)0.3

18
Although the specimens were tested using this high value of concrete strength, all specimens
showed very high displacement capacities and satisfied the 6 mm criterion of EN 1994-1-1. The
influence of concrete relaxation (see load drops in the load slip curve) on the total shear resistance
was about 10%.

Typical behaviour with Cofraplus 60 decking


Specimens with Cofraplus 60 decking (known as CP60 in this report), the load-slip curves showed
the development of two peaks, as shown in Figure 2.14. After the first load peak had occurred, rib
punch through failure was observed. For further loading, the test load increased as tension forces
developed in the stud shank, because of the restraint to displacement of the head of the stud. A
second peak load was reached and finally failure occurred as concrete pull-out, shown in Figure
2.15, or stud failure.

Figure 2.14 Typical load-slip curve for specimen with CP 60 decking (specimen 1-04-1)

Typically the studs developed two yield hinges, as shown in Figure 2.15. The observation of rib
punch-through failure after the first load peak shows, that at this point the plastic hinges had
developed in the stud shank. Therefore, the first load peak is characterised by the plastic
displacement of the stud.

Figure 2.15 Failure surface for concrete pull- Figure 2.16 Stud deformation
out for CP60 decking with 2 plastic hinges for CP60
decking
All specimens with CP 60 decking developed very high displacements and satisfied the 6 mm
criterion of EN 1994-1-1. The major influences on the load-slip behaviour of tests with CP 60
decking were:

 The resistance of pairs of shear connectors was approximately twice that for single shear
connectors, which showed that the mode of failure was not dominated by development of a
concrete ‘cone’ over the shear connectors.
 The shear resistance for 22 mm diameter shear connectors was 65% more than for 19 mm
diameter shear connectors.

19
 Higher concrete strengths lead to higher resistances but also reduced the displacement
capacity.
Eccentric transverse loading also did not affect the first load peak significantly. The second peak
load decreased, while the displacement capacity increased. It is assumed that cracking of the
concrete weakened the embedment conditions of the head of the stud.

Typical behaviour with ComFlor 80


The specimens with ComFlor 80 decking (known as CF 80 in this report) developed a typical load-
slip behaviour as shown in Figure 2.17. The observed failure mode was rib pry-out failure. This is
the failure of a concrete cone at low slip of about 1.0 to 4.9 mm (see Figure 2.17). The stud
deformation typically showed only one yield hinge for this deck type, as shown in Figure 2.18.

After rib pry-out failure had occurred, the load bearing mechanism changed into a mode that
depends on frictional resistance and aggregate interlock over the failure surface. This mode was
sensitive to transverse loading. The change of the load bearing mechanism typically was
accompanied by a temporarily reduction in the test load. In some cases, the test load fell below the
characteristic resistance PRk, which may lead to low values of the displacement capacity δu. In most
cases, the test load increased later in the loading when the load bearing mechanism had changed
into the frictional mode. Then the displacement behaviour was generally very ductile.

Figure 2.17 Typical load-slip curve for specimen with CF80 decking (specimen 1-10-3)

Figure 2.18 Failure surface for rib pry-out Figure 2.19 Stud deformation with one plastic
failure for CF 80 decking hinge for CF 80 decking

Typical behaviour with Cofrastra 56


A second series of push-out tests was conducted using re-entrant decking Cofrastra 56 of 56 mm
depth (known as CF56 in this report). The rib height is almost the same as for the CP 60 profile
(56 mm versus 58 mm). These specimens had the load-slip behaviour shown in Figure 2.20.
Because of the narrow rib width, it was difficult to observe the failure mode and the only observed
failure mode (noted from outside) was rib punch-through failure.

20
Load-slip curve 3-03 P1
600

500

384,0 kN
400
Load [kN]

300 381,5 kN

200

100
5,80 mm

0
0 5 10 15 20 25 30 35 40
Slip [mm]
Figure 2.20 Typical load-slip curve for specimen with CF56 decking (specimen 3-03-1)

Typical behaviour with the decking parallel to the beam


Series KL investigated the load-slip behaviour when the decking is placed parallel to the steel
beam. The load-slip curves obtained for this series are shown in Figure 2.21. At a slip of about 4 to
6mm, cracks at the top of the ribs were observed. For all three specimens, rib shearing was
identified as the failure mode. The shear connectors showed plastic bending deformation with
single curvature.

The maximum test load was between 330 and 354 kN. According to EN 1990 Annex D.7.2, a
characteristic resistance of 296kN was determined which is about 90% of the minimum test
resistance. The factor kn=3.37 was determined for the case “Vx unknown”. The characteristic
resistance lead to slip capacities of 8.6 to 9.2mm. The characteristic slip capacity is 7.7mm, which
means that the shear connection was ductile.

Figure 2.21 Load-slip curves for specimens with parallel deck ribs in series KL

Evaluation of push test results was made to EN 1994-1-1 Annex B-2. Table 2.4 summarises the
results of the push-out tests. The resistances per shear stud, Pe do not consider relaxation and the
static resistance per shear stud Pe,static considers relaxation. Based on the values Pe, the
characteristic resistance PRk of a series was taken as the minimum resistance in the series reduced
by 10%. The same procedure was applied to obtain the static value of the characteristic resistance
PRk,static, by evaluating the test results Pe,static.

21
Table 2.4 Push-out test results for CF 80 decking in Task 1.2
Series i Deck d nr fc Pe Pe,static PRk PRk,static u‡ δuk‡
[mm] [N/mm²] [kN/stud] [kN/stud] [mm] [mm]
1 - 19 - 44.2 157.94 143.63 11.16
1-01 2 - 19 - 44.3 169.00 153.25 142.15 129.27 13.43 10.04
3 - 19 - 44.8 168.69 151.06 11.20
1 - 22 - 45.6 215.31 195.31 10.56
1-02-1 2 - 22 - 45.5 216.75 196.88 193.78 175.78 13.69 9.50
3 - 22 - 44.2 218.43 198.50 13.38
1 - 22 - 46.2 206.81 188.56 11.58
1-02-2 2 - 22 - 45.8 211.56 189.13 186.13 169.09 14.47 10.42
3 - 22 - 45.4 208.68 187.88 14.11
1 CP60 22 1 41.0 81.50 74.63 46.70
1-03 2 CP60 22 1 42.5 93.88 86.00 73.35 67.16 40.70 36.63
3 CP60 22 1 42.9 97.13 88.88 45.40
1 CP60 22 1 30.1 73.91 70.44 67.62
1-04 2 CP60 22 1 30.9 73.61 69.63 61.79 58.53 61.48 55.33
3 CP60 22 1 30.9 68.66 65.06 68.65
1 CP60 22 1 31.7 69.95 66.48 62.08
1-05 2 CP60 22 1 31.7 79.15 75.23 62.96 59.76 61.64 51.62
3 CP60 22 1 32.6 74.43 70.68 57.36
1 CP60 22 1 29.9 73.11 68.89 71.07
63.12 59.12 59.37
1-06 2 CP60 22 1 31.1 70.13 65.69 65.97
3 CP60 22 1 44.0 90.31 83.44 81.28 75.10 41.37 37.23
1 CP60 22 1 45.1 89.25 83.81 58.73
1-07 2 CP60 22 1 40.9 92.77 86.52 67.29 63.19 68.03 52.86
3 CP60 22 1 42.6 74.77 70.12 67.42
3.26
1 CF80 19 2† 42.2 45.52 42.80
23.21↔ 2.71
40.97 38.52
1-08 3.01 5.54↔
2 CF80 19 2† 42.2 46.09 43.33
6.15↔
3 CF80 19 2† 42.2 48.26 45.92 43.43 41.33 25.27 22.74
1 CF80 19 2† 42.6 42.28 39.11 38.05 39.11 7.59 6.83
1-09 2 CF80 19 2† 42.6 50.32 46.40 -
43.37 41.35 -
3 CF80 19 2† 42.6 48.19 45.98 -
1 CF80 19 2† 42.6 45.64 40.57 41.08 36.68 25.74 23.17
5.11 4.60
2 CF80 19 2† 42.6 46.44 39.29 41.80 35.36
1-10 9.05↔ 8.15↔
5.04 4.54
3 CF80 19 2† 42.6 35.46 31.22 31.91 28.10
8.97↔ 8.07↔
1 CF80 19 2† 42.6 54.64 51.41 16.86
1-11 2 CF80 19 2† 42.6 52.68 46.04 44.40 41.14 20.52 15.17
3 CF80 19 2† 42.6 49.33 45.71 22.32
3.68 3.31
1 CF80 19 1† 44.1 79.07 71.84 71.17 64.67
8.22↔ 7.40↔
NR1 5.00 4.50
2 CF80 19 1† 45.7 74.93 67.93 67.49 61.13
13.56↔ 12.18↔
3 CF80 19 1† 44.7 70.46 64.87 63.41 58.38 3.27 2.94
‡ Determined from PRk
† Through deck welded

↔ Second intersection of P-δ-Curve and P


Rk

22
The displacement capacity δu was determined according to EN 1994-1-1 Annex B2 using the values
for PRk. The displacement capacity δu then is the slip, at which the test load drops for the first time
below PRk. The characteristic slip capacity δuk then is the minimum value for δu reduced by 10%.

As mentioned above, tests with CF80 decking exhibited rib pry-out failure after which the load
bearing mechanism changed to a frictional mode. The change of the load bearing mechanism was
accompanied by a reduction of the test load. In some cases, this reduction fell below PRk and led to
readings for δu of less than 6mm. As the load-slip behaviour generally was very ductile after the
load bearing mechanism had changed, the second intersection δu2 between PRk and the load-slip
curve was also evaluated where appropriate. Typically the values for δu2 satisfy the 6mm criterion.

2.4 Influencing parameters

2.4.1 Influence of boundary conditions


When tension ties were used, no forces in the tension ties were observed as the frictional
resistances at the supports appeared to be sufficient to resist any horizontal forces that developed.
For specimens with CF80 decking, it was observed that the slabs were subjected to bending during
testing. Because of this, significant forces developed in the tension ties and strongly affected the
load-slip behaviour.

The results give evidence that constant transverse loading can be applied by controlling this
rigidity. This was achieved using non-linear type C disc springs according to DIN 2093, which would
render the use of additional hydraulic actuators unnecessary.

Figure 2.22 Load-slip curves with and without tension ties for specimens using CF80 decking

23
Table 2.5 Push-out test results for CF 80 decking in Task 1.2

Series i Deck d nr fc Pe Pe,static PRk PRk,static u‡ δuk‡


[mm] [N/mm²] [kN/stud] [kN/stud] [mm]
1 CP60 19 1† 42.4 65.25 57.88 -
2-01 2 CP60 19 1† 42.6 70.13 62.75 58.73 52.09 - -
3 CP60 19 1† 41.8 74.88 66.63 -
2-02 1 CP60 19 1† 40.7 69.88 62.00 62.89 55.80 - -
2-03 1 CP60 19 1 39.7 57.63 51.63 51.86 46.46 - -
2-04 1 CP60 19 2† 40.3 68.38 61.50 61.54 55.35 7.39 6.65
1 CP60 19 2† 40.2 65.63 57.94 9.71
2-05 2 CP60 19 2† 38.6 71.06 64.44 59.07 52.15 - 8.37
3 CP60 19 2† 39.2 67.06 59.50 9.30
1 CP60 22 1 40.1 87.50 79.13 -
2-06 2 CP60 22 1 39.6 101.60 83.38 78.75 71.22 23.70 21.33
3 CP60 22 1 40.6 96.50 86.75 -
1 CP60 22 1 46.4 114.60 97.25 8.92
2-07 2 CP60 22 1 46.3 112.40 100.38 96.84 87.53 7.83 7.05
3 CP60 22 1 46.0 107.60 99.75 27.93
2-08 1 CP60 19 1 45.8 59.10 51.00 53.19 45.90 - -
1 CF80 19 2† 46.0 47.31 38.86 32.88
42.58 34.97 29.59
3-01 2 CF80 19 2† 46.0 54.92 51.66 35.31
3 CF80 19 2† 40.4 52.78 47.32 47.50 42.59 2.39 2.15
3-02 1 CF80 19 2† 42.6 37.00 34.90 33.30 31.41 1.80 1.62
3-03 1 CS56 19 1† 39.9 96.00 84.63 86.40 76.17 - -
3-04 1 CS56 19 1 41.0 112.90 95.63 101.61 86.06 - -
3-05 1 CS56 19 2† 39.7 95.10 87.19 85.59 78.47 7.76 6.98
3-06 1 CS56 19 2† 39.5 94.60 85.13 85.14 76.61 8.19 7.37
3-07 1 CS56 19 2 40.7 103.90 93.88 93.51 84.49 - -
3-08 1 CS56 19 1 41.3 109.10 97.25 98.19 87.53 - -
3-09 1 CS56 22 1 40.7 128.50 117.13 115.65 105.41 5.99 5.39
3-10 1 CS56 22 1 40.5 121.25 107.50 109.13 96.75 - -
‡ Determined from PRk
† Through deck welded

2.4.2 Influence of concentric transverse loading


For specimens using CP 60 decking, the application of transverse loading led to only a small
increase of the shear resistance but the displacement capacity was slightly reduced. It was
observed that the natural variability of the resistance was larger than the beneficial influence of
transverse loading. Thus, it can be concluded that transverse loading was not significant for these
tests failing by S-shaped bending of the shear connectors.

24
Figure 2.23 Influence of concentric transverse loading using CP 60 decking

Figure 2.23 shows the load-slip curves obtained with CF 80 decking and pairs of studs per rib for
different degrees of transverse loading. It was observed that transverse loading improved the
failure load of the ribs. The load-slip curves obtained with single studs per rib are shown in Figure
2.24. There, the test load was decreased for higher transverse loading. However, up to a slip of
about 18mm, the variation of the test results appears to be within a reasonable variability for
concrete failure and frictional resistances.

Figure 2.24 Influence of concentric transverse loading for CF 80 with pairs of studs per rib

Figure 2.25 Influence of concentric transverse loading using CF80 and single studs per rib

25
2.4.3 Influence of eccentric transverse loading
As an example, Figure 2.26 shows the influence of eccentric transverse loading when CP 60
decking was used. As for concentric transverse loading, the resistance was not significantly
influenced, but the displacement capacity increased.

Figure 2.26 Influence of eccentric transverse loading using CP 60 decking

Figure 2.27 shows a comparison of the load slip-curves for specimens with CF 80 decking with a
concentric and a constant eccentric transverse load of 17.5kN per side. The load-slip behaviour is
in principle the same, but for tests with eccentric transverse loading, higher failure loads were
identified. The displacement capacity with eccentric transverse loading was also significantly higher
than for the reference test without transverse load.

Figure 2.27 Influence of eccentric loading using CF80 decking

2.4.4 Influence of re-entrant deck shape


The trapezoidal (CP60) and re-entrant (CP56) decking profiles were almost the same rib height.
However, the shear resistance of the re-entrant deck slabs in series 3-03 was about 37% more
than the same tests using CP60. This is because of the clamping effect of the re-entrant profile
which holds the concrete in the rib and causes an additional resistance, as shown in Figure 2.28.

26
Influence of the type of profiled sheeting
400

350

300

250
Load [kN]

200

150

100

50

0
0 2 4 6 8 10
Slip [mm]

Series 2-01-1 Cofraplus60 Series 3-03-1 Cofrastra56

Figure 2.28 Influence of the deck shape (comparison of tests 2-01-1 and 3-03-1)

2.4.5 Influence of the number of shear studs


If shear connectors are placed in pairs per deck rib, there is an overlap in the zone of action of
each stud meaning that the total resistance will be less than the resistance of a single stud.
EN 1994-1-1 allows use of two shear connectors per rib if their spacing is more than 4dsc, where dsc
is the shear connector diameter. The maximum load per stud is reduced by the factor 1/√nr=0.71.
Konrad increased the factor from 0.71 to 0.80 because of his test results and numerical studies,
but this applies when the mode of failure is not due to a concrete ‘cone’ over the shear connectors.

The performed tests in this research had a centre distance of 100 mm between the two shear
connectors per rib, which is approximately 5dsc for dsc= 19 mm diameter shear connectors. The
maximum test load per specimen using two shear connectors per rib achieved nearly twice the load
that was reached in the comparison to those with single shear connectors.

For specimens with CF80 decking this was not the case. In these tests, the failure mode was by rib
pry-out and therefore the test load was governed by concrete failure. Without transverse loading,
the comparison of test NR1-1 with single studs per rib to test 1-10-3 with pairs of studs per rib
shows only a difference of 10% in the failure load. The difference of the load-slip behaviour was
very similar until at about 19 mm, stud failure occurred for specimen NR1-1. The similar load-slip
curves are because the behaviour is governed by the size and shape of the failure cone, which did
not differ much from each other, and by the concrete strength and coefficient of friction.

Figure 2.29 Influence of number of studs per rib for rib pry-out failure (comparison of tests 1-10-
3 (pairs of studs) and NR1-1 (single studs))

2.4.6 Influence of the stud diameter


According to EN 1994-1-1, the failure load per shear stud is in proportion to dsc2. The performed
tests used shear studs of 19 mm or 22 mm diameter. According to the EN 1994-1-1, the failure
load should increase by (22/19)² = +34%. Figure 2.30 shows that the first load peak of specimen

27
2-06-3 with 22 mm diameter shear studs is about 65% higher than for specimen 2-03-1, nearly
twice as high as the prediction.

According to Lungershausen, the resistance depends on the bending resistance of the stud shank.
The influence of the stud diameter then is (22/19)³ = +55%, which is much closer to the observed
influence.

Influence of the diameter of the shear connectors


400

350

300

250
Load [kN]

200

150

100

50

0
0 2 4 6 8 10
Slip [mm]

2-03 19mm shear connectors 2-06 3 (22mm shear connectors)

Figure 2.30 Influence of the diameter of the shear connectors

2.4.7 Influence of the welding procedure


A total of 11 of 27 push-out tests used decking with through-deck welded shear studs. They usually
have a higher bearing capacity because part of the shear force is dissipated directly by the decking
and additional concrete compression struts. Therefore, it was expected that all configurations with
through-deck welded shear studs would show a higher bearing capacity than those with pre-
punched decks.

For decks such as CP60 (t=0.88mm), EN 1994-1-1 and the former DIN 18800-5 take an increase
between 13% and 17% compared with profiled decking with cut holes. An average increase of the
failure load of 22% was observed when using through-deck welding in combination with CP 60 (see
Figure 2.31 where results from specimen P1 of series 2-01 with pre-punched decking and specimen
P1 of series 2-03 with through-deck welded studs are compared). However, this effect was
reversed when re-entrant decking was used, which was not expected. The results showed a
decrease of the failure load of 10%-18% for through-deck welded configurations. There is no
explanation for this effect.

Influence of stud welding method


300

+22%
250

200
Load [kN]

150

100

50

0
0 2 4 6 8 10
Slip [mm]

2-01 P1 (pre-punched) 2-03 (Through deck welding)

Figure 2.31 Influence of stud welding procedure using CP 60 (tests 2-01-1 and 2-03-1)

28
Figure 2.32 Influence of the stud welding procedure using CF 80 (tests 1-10-3 (through-deck
welded) and 3-02 (pre-punched deck))

2.4.8 Influence of the reinforcement pattern


Current standards do not provide any information with regard to using one or two layers of
reinforcement for composite beams using profiled decking. The performed push-out tests
investigated the influence of a second layer of reinforcement. All specimens with CP60 and CP56
decking had a Q335 (335mm²/m) mesh on top of the concrete slab. In addition, some specimens
had a Q188 (188mm²/m) mesh that was positioned directly above the rib of the profiled decking
(see Figure 2.33).

The test results showed that two layers of reinforcement increase the ductility by a large amount,
which means that the slip of the first maximum peak increased by almost a factor of two However,
the shear resistance was not influenced significantly- compare specimen 2-01-1 with one
reinforcement layer to specimen 2-02-1 with two reinforcement layers.

Figure 2.33 Influence of reinforcement pattern (comparison of one and two layers)

For tests with CF80 decking, where rib pry-out failure was observed, the influence of the
reinforcement was significant. When one layer of reinforcement was used, it was placed 30 mm
above the top of the rib and therefore just below the head of the stud. For two layers, it was
necessary to place the bottom layer directly on the decking. Because of the stiffeners on top of
CF80 decking, the bottom layer therefore was placed 15 mm above the rib, i.e. only 15 mm lower
than for a single layer of reinforcement.

The failure load for rib pry-out increased by 49% when two layers of reinforcement were used. It is
assumed that the bottom wire fabric reinforced the top part of the failure cone around the head of
the stud. As soon as the frictional load-bearing mechanism was established, the reinforcement had
no influence on the test load.

29
Figure 2.34 Influence of reinforcement pattern on rib failure (tests 1-10-3 and 3-01-3))

2.4.9 Influence of concrete strength


In compliance with EN 1994-1-1, the design failure load PRd per shear connector for concrete failure
with transverse decking is in proportion to √fck, as follows:

1
𝑃𝑅𝑑 = 0.29 ∙ 𝑑2 ∙ √𝑓𝑐𝑘 ∙ 𝐸𝑐𝑚 ∙ (2.2)
𝛾𝑉

The studies of Konrad could not confirm the correlation of the failure load with √fck. In case of
concrete failure, his numerical studies resulted in the failure load being a factor of fck2/3, as follows:

2
𝑓𝑐𝑘 3 𝑓𝑢𝑘 1
𝑃𝑅𝑑 = [313 ∙ 𝐴𝑊𝑢𝑙𝑠𝑡,𝑒𝑓𝑓 ∙ ( ) + 250 ∙ ( ) ∙ 𝑑2 ] ∙ (2.3)
𝑓𝑐,𝑛𝑜𝑚 𝑓𝑢,𝑛𝑜𝑚 𝛾𝑉

The rules of EN 1994-1-1 seem to be restrictive compared to the new test results from this study.
The comparison of the two series showed that the measured concrete strength for specimen 2-06-3
was 40N/mm² and for specimen 2-07-3, it was 46 N/mm², which is an increase of 15% in concrete
strength. The first load peak increased from 386 kN up to 431 kN, which is an increase of 11.5%.
EN 1994-1-1 would only allow an improvement of 7% while the formula by Konrad gives an
increase of 10%, which fits much better to the test results.

Figure 2.35 Influence of concrete strength on shear connector resistance

As transverse loading was not significant for specimens with CP 60 decking, test series 1-04 can
also be considered. There, the measured concrete strength was about 30N/mm², which is 25% less
than for specimen 2-06-3. The resistance Pe1 was on average 25% less. However, for the tests in

30
this comparison, a significant contribution of the stud itself to the shear resistance is assumed,
because of its S-shaped bending deformation.

2.5 Comparison with characteristic resistances to EN 1994-1-1


According to EN 1994-1-1, the characteristic resistance of a headed shear stud in the rib of a
composite slab is calculated by multiplication of the resistance in a solid slab with the factor kt. The
reduction factor kt is calculated according to equation (4.8). The upper limit kt,max is given in
equation (2.4):

0.70 𝑏𝑚 ℎ𝑠𝑐
𝑘𝑡 = ( − 1) ≤ 𝑘𝑡,𝑚𝑎𝑥 (2.4)
√𝑛𝑟 ℎ𝑝 ℎ𝑝

The characteristic resistance of headed studs in the ribs of composite slabs placed transverse to the
beam is then calculated according to equation (2.5).

𝑃𝑅𝑘,𝑐 = 0.29 𝛼 𝑑2 √𝑓𝑐𝑘 𝐸𝑐𝑚


𝑃𝑅𝑘 = 𝑘𝑡 min { 2 (2.5)
𝑃𝑅𝑘,𝑠 = 0.80 𝑓𝑢 𝜋 𝑑 ⁄4

The obtained results are compared to the static characteristic test resistances Pe,k, which are
assumed to be according to equation (2.6).

𝑃𝑒,𝑘 = 0.9 𝑃𝑒,𝑠𝑡𝑎𝑡𝑖𝑐 (2.6)

This simplified assumption allows the comparison to be made of each test result. Therefore more
information on the coefficients of variation and correlation can be obtained.

For CF 80 decking, the results are un-conservative by about 16%. These were the only tests with
wide ribs, where the geometric term of equation (2.4) was important. For specimens with CP 60
decking, the reduction factor was at the limit, kt,max . The results are un-conservative by 26% to
34%, but show a good coefficient of variation. In case of CP 56, the reduction factor was at the
limit, kt,max. For single studs, the results are un-conservative by 6% and for pairs of studs they are
conservative by 18%.

In the second step, characteristic material properties and limitations given in EN 1994-1-1 were
considered. The concrete strength was taken according to equation (2.7). The steel strength is
taken according to equation (2.8) with the limit of 450N/mm² according to EN 1994-1-1. The factor
𝑘∞ is taken as 1.64 and the coefficient of variation Vfu as 0.05.

𝑓𝑐𝑘 = 𝑓𝑐𝑚 − 8 [N/mm²] (2.7)

𝑓𝑢𝑘 = 𝑓𝑢 (1 − 𝑘∞ 𝑉𝑓𝑢 ) ≤ 450 N/mm² (2.8)

Using the characteristic material properties and considering their limits in EN 1994-1-1, the
predicted resistances decrease by about 10%, but the composite slab tests reach on average only
88% of the predicted resistance (see Figure 2.36).

For all specimens with CP60 and CP 56 decking, the upper limit kt,max controls, whereas for CF 80
decking, the reduction factor kt is based on the geometry according to equation (2.4). For
specimens with CF 80 decking, the formulae predicts resistances that are relatively close to the
test results. For about two-thirds of the tests, the shear resistance was limited by the steel
resistance in equation (2.5).

In the case of CP 60 with single studs per rib, equation (2.4) results in kt values of about 1.10 and
for pairs of studs in values of about 0.75. In all cases, the limit on kt,max controls. Nevertheless, a
good coefficient of correlation is obtained, but the resistance is over-estimated by 20% to 26%.

31
200

150 CF80-nr=1
Pe,k [KN/stud] CF80-nr=2
100 CP60-nr=1
CP60-nr=2
50 CS56-nr=1
solid slab
0 CS56-nr=2
0 50 100 150 200
PRk [kN/stud]

Figure 2.36 Comparison of "characteristic" test results to predicted characteristic resistance for
characteristic material properties

2.6 Recommended standard test regime


EN 1994-1-1: Eurocode 4 currently does not include rules specifically developed for push-out tests
when relatively deep steel decking is placed transversely to the beam. The following sections give
guidelines for the design of the specimen geometry and the conduction of the push-out test.

 Considering the geometry of the test specimen, the height of the concrete slab must be a
function of the decking and cannot be fixed to a certain dimension. Only a general guideline for
the shape of the composite slab can be given.
 The conducted push-out tests showed that in most tests the failure of a concrete cone occurred.
The width of the specimen should therefore be chosen to fully incorporate this cone. Therefore,
the width of the specimen is a function of the angle of the failure plane and the height of the
studs. Recommendations for the width of the specimen can be based on a conservative
assumption for the angle of the failure planes.
 The following guidelines are suggested:
 the depth d, of the slab should reflect typical applications;
 the minimum width, b, of the concrete slabs should be chosen according to Figure 2.37 and
should be within h≤b≤1.5h;
 the height, h, of the specimen should be chosen according to the dimensions of the decking,
matching in general the shape as given in Figure 2.37;
 the recess at the bottom of the specimen according to EN 1994-1-1 is optional. Sufficient
transverse reinforcement to prevent vertical splitting of the slab must also be placed above
the recess. The required reinforcement can be determined using a truss model for the
distribution of forces (see Figure 2.38):
 the specimen should be placed in a mortar bed on the supporting faces;
 the specimen should be placed on a layer of Teflon which is placed between steel plates to
allow the horizontal displacement of the slabs due to the applied loads and prevent frictional
resistances from affecting the distribution of the transverse loading;
The conducted push-out tests showed that the influence and relevance of transverse loading
depends on the failure modes which occur.

hsc

25°

≥100 hsc/tan25° es hsc/tan25° ≥100

Figure 2.37 Minimum width of concrete slab for push-out tests

32
With: cot=1.2  ≈40°

a h≥3a
H

Recess is optional
 

~ A/3 ≤ C ≤ 2/3·A
7  Reaction force to be
30 within the core of the
A 200
supporting face
H ≤ B ≤ 1.5H

Length of test specimen Strut and tie model for equilibrium at recess

Figure 2.38 Dimensional and detailing requirements for push-out test

2.7 New mechanical model for shear connector resistance based on this
research
In the scope of this project, a new mechanical model was developed based on equilibrium of a
shear connector embedded in concrete with transverse decking based on the thesis of Nellinger
(2015). It was observed that the model using the plastic bending resistance of shear connectors to
predict its shear resistance showed a better correlation with tests than the empirical formulae in
Eurocode 4. In this research the influence of the concrete strength and the shear connector
position was considered.

2.7.1 Proposed shear connector resistance formulae


This method is based on the elastic bending resistance of a concrete failure ‘cone’ over the shear
connector acting as a cantilever. The required cross-section properties were derived from the
observed concrete failure surface in push-out tests. The influence of the welding position, such as a
favourable position, was then considered as an eccentric compression force acting on the concrete
rib. The empirical derived spacing of the yield hinges, which did not match the observed spacing in
tests, was replaced with a value obtained from the properties of the concrete failure surface.
Furthermore, the observed single curvature of the shear connector for rib pry-out failure and
double curvature of the stud for rib punch-through failure were considered.

These considerations have led to a proposed new equation (2.9) for shear connector resistance as
influenced by concrete failure. The upper bound for the shear resistance is given by the plastic
resistance of the steel shank in pure shear, given by equation (2.10), which is 30% lower than in
equation (2.5). The design shear resistance is the smaller value of these two equations. This also
includes the beneficial effect of out of plane load acting on the slab.

𝑁𝑞 + 𝑁𝑠𝑐
(𝛼𝑐𝑡 𝑓𝑐𝑡𝑚 + ) 𝑊 + 𝑁𝑠𝑐 𝑒𝐿 𝑛𝑦 𝑀𝑝𝑙 1
𝐴
𝑃𝑅𝑑 = 0.84 [ + ] (2.9)
ℎ𝑝 𝑛𝑟 ℎ𝑠 − 𝑑/2 𝛾𝑉

𝑑2 1
𝑃𝑅𝑑 = 0.56 𝑓𝑢𝑘 𝜋
4 𝛾𝑉 (2.10)

Where: 𝛾𝑉 = 1.25 Partial safety factor for shear connector resistance

𝛼𝑐𝑡 = 0.85 Influence of relaxation on concrete strength

𝑁𝑞 Transverse load applied per deck rib

𝑑2 Compression of concrete rib at the head of the stud


𝑁𝑠𝑐 = 0.1 𝑛𝑟 𝑓𝑢𝑘 𝜋
4

𝐴 = (2.4 ℎ𝑠𝑐 + ∑ 𝑒𝑠 ) 𝑏𝑚𝑎𝑥 Effective area of concrete failure cone

33
3
1 𝑏𝑚𝑎𝑥
𝑊 = (2.4 ℎ𝑠𝑐 + ∑ 𝑒𝑠 ) Section modulus of concrete failure cone
6 𝑏𝑜

(ℎ𝑝 + ℎ𝑠𝑐 ) ∑ 𝑒𝑠
𝛽 ℎ𝑠𝑐 +
4.8 ℎ𝑠𝑐 Position of upper yield hinge in the shear connector
ℎ𝑠 =
∑ 𝑒𝑠
1+
2.4 ℎ𝑠𝑐
ℎ𝑠𝑐 − ℎ𝑝
1 ≤ 𝑛𝑦 = 2 ( − 1) ≤ 2 Number of yield hinges in the shear connector
𝑑 √ 𝑛𝑟
1
𝑀𝑝𝑙 = 𝑓𝑢𝑘 𝑑3 Plastic bending resistance of shear connector
6
0.45 trapezoidal decking Factor depending on the deck shape
𝛽={
0.41 re-entrant decking

Other geometric parameters are presented in Figure 2.39.

Figure 2.39 Geometric parameters used in the new model

The use of the above formulae has the following limitations:

 Tensile strength of the studs: fuk ≤ 500 N/mm²


 Concrete strength: 20 ≤ fck ≤ 50 N/mm²
 Rib height: hp ≤ 155 mm
 Embedment of the head of the stud: hsc – hp ≥ d
 Stud height: hsc ≥ 4dsc
 Stud diameter: dsc ≥ 16 mm and for through deck welded studs d ≤ 19 mm and when holes
are cut in the decking dsc ≤ 22mm
 Studs per deck rib: nr ≤ 3
 Studs should ideally be welded at the centreline of the rib. If this is not possible, they should
be welded in a ‘staggered’ or ‘favourable’ position.

2.7.2 Comparison of the new formula to test results


The results of the statistical analysis using all collected push tests results are shown in Figure 2.40.
Because this database considers cases that are not covered by the limitations of the formulae, the
design resistance has been reduced by 18% according to EN 1994-1-1. The coefficient of
correlation R = 0.73 is also better than the previous results: R = 0.56 for EN 1994-1-1, R = 0.65
for Konrad’s complex solution and R = 0.55 for Konrad’s simple solution. The coefficient of
variation V = 0.28 is significantly better than V = 0.41 for EN 1994-1-1, but similar to V = 0.25 for
Konrad’s complex solution and V = 0.27 for Konrad’s simple solution.

34
Figure 2.40 Statistical analysis of the model of Nellinger for 375 push- tests

Considering the limitations for the above formulae, the results shown in Figure 2.41 are obtained.
The coefficient of variation improves to V = 0.18, which is smaller than for the other models in
their field of application. Likewise, the highest coefficient of correlation R = 0.85 is obtained. In
addition, the predicted design shear resistance differs by only 1% from the requirement of EN
1990. Furthermore, the range of application is wider.

Figure 2.41 Statistical analysis of the model of Nellinger in its range of application

2.8 Comparison of initially planned activities and work accomplished


The push test series on the shear connectors used in conjunction with the three deck profiles was
performed with the following additional tests:
 CF 80 decking with single shear connectors and with the influence of transverse loading in order
to provide data for use in interpreting the beam tests in WP3.

35
 CP 60 decking with decking orientated parallel to the steel profile with single shear connectors
welded directly to the beam at 200mm spacing and without the influence of transverse loading
in order to provide data for use in interpreting the floor plate tests in WP4.

2.8.1 Description and impact of the research results


For specimens with CP 60 decking which had narrow ribs and a large embedment depth of the
studs, the mode of failure was dominated by bending of the shear connectors and so the influence
of transverse loading was not important.

For specimens with CF 80 decking, which had wide ribs and a relatively small embedment depth,
the load-slip behaviour was affected by concrete ‘cone’ failure In this case, the influence of
transverse loading was noticeable. This leads to the following recommendations:

 If the height, hsc, of the studs after welding >1.56hp, transverse loading does not need to be
applied. If hsc ≤ 1.56 hp, a transverse load of 10% of the shear force per side should be
applied. The transverse load can be applied as a constant value during the tests
 It is recommended to apply the transverse load using a hydraulic jack. Other methods of load
application (e.g. pre-stressed tension ties with type C disc springs according to DIN 2093) are
allowed, if a horizontal displacement of at least 15mm on each side can be guaranteed without
increasing the transverse load by more than 10% above the target value.
 The consideration of an eccentricity for transverse loading is optional. If the transverse load is
applied eccentrically, the load must be applied outside of the failure cone shown in Figure 2.37
and should be chosen to reflect the bending moment diagram as realistically as possible. The
minimum eccentricity is 380 mm.

2.8.2 Concluding remarks for WP1


The tests have shown that there are two distinct modes of failure which are dependent on either a
concrete ‘cone type’ failure or S-shaped bending of the shear connectors. The transition between
the two modes depends on the embedment length of the shear connectors in the concrete topping
as a proportion of their height.

The tests have shown the importance of transverse loading for cases in which the shear connectors
fail by a mode of concrete failure. In this case, there is a small increase in failure load but a larger
increase in deformation capacity. Transverse loading is not important where the shear connectors
fail by bending. A method of test in which constant transverse load is applied is proposed.

2.8.3 Exploitation and impact of the research results


The tests were important in defining the two modes of failure, and the choice of the two deck
profiles proved to be important in this respect. This is linked to the study on shear connector
resistances in WP5.1. The new method of test is an important outcome of this research, which will
be fed into Eurocode 4.

The new formulae for the shear connector resistances are being considered in the development of
Eurocode 4 and will be subject to further calibration work and ‘ease of use’ tests, so may be
modified subsequently.

36
3 WP2: Composite beam tests to investigate the influence of low degrees of
shear connection

3.1 Objectives of WP2


The objective of WP2 was to set up a series of beam tests with low degrees of shear connection
(less than 40%) in which parts of the beams were subject to constant shear. Two series of tests
were performed as foreseen in the proposal:

 Tests on 5 m span IPE 300 beams with CP 60 decking (56 mm height):


 Tests on 6 m span IPE 360 beams with CF80 decking (80 mm height):
The results are compared with the plastic bending resistances for partial shear connection using the
shear connector resistances obtained from push-tests in WP1. The comparison is made in the form
of a ratio of test bending resistance to the calculated bending resistance, which is then used to
determine the suitability of the partial shear connection rules in Eurocode 4.

3.1.1 Comparison of initially planned activities and work accomplished


A series of 10 beam tests (8 at USTUTT, 2 at ULUX) was conducted to provide a comparison of the
relative performance for the effects of the shear connector diameter, the use of through deck
welding or use of pre-punched holes in the decking, and single or pairs of shear connectors per
deck rib. The calculated degree of shear connection based on design properties was estimated to
be in the range of 20 to 40% (actually it was 15 to 42% based on the tests in WP1).

In accordance with the proposal, the composite beam tests covered the following parameters:

 Shear connectors (19 mm diameter) welded through the profiled sheeting or welded directly to
the beam through punched holes in the decking
 Shear connectors (22 mm diameter) welded directly to the beam
 Number of shear connectors per deck rib (single or pairs)
 Different deck profiles (CF80 and CP60 (58 mm deep))
 Influence of reinforcement in one or two layers
 Influence of concrete strength (C30/37 or C40/50)
The different spans used for the tests result from the different deck profiles that were used in the
tests. The planned tests in the proposal were for the same span for all deck profiles. However, the
WP partners agreed that the geometry of the beam and the size of the profiled decking must be
compatible. SCI proposed that the span/depth ratio of the beam should be close to 20 to ensure
that the deformation characteristics are typical of the normal range of application.

This lead to the decision that University of Stuttgart carried out the tests with a 5 m span IPE 300
beam with CP60 decking and University of Luxembourg carried out the tests using a 6m span
IPE 360 beam with CF 80 decking. The corresponding effective widths beff therefore were 1.25 m
and 1.50 m respectively. The detailed configuration of the beam tests is given in Table 3.1.

To ensure that there was uniform shear over the outer parts of the composite beam, the load
introduction points were moved closer to mid-span than the third points foreseen in the proposal.
Therefore, the shear span was 1.80 m for the 5 m tests and 2.25 m for the 6 m tests (see Figure
3.1 and Figure 3.2). The University of Stuttgart applied the point loads centered on the steel beam
while University of Luxembourg applied the loads distributed over the effective width of the beam
using a load distribution system.

37
Table 3.1 Beam tests and their parameters performed in Task 1.3

Single
Beam, slab and deck Shear connector Concrete Reinforcement
Series /pairs
configuration diameter and welding grade layers
per rib
1 layer
2-01 19 mm through-welded
Single
2-02 19 mm punched hole 2 layers

IPE300 beam, 5 m span C30/37


2-03 19 mm through-welded 1 layer
with130 mm deep slab Pairs
2-04 comprising 19 mm punched hole 1 layer
2-05 CP60 decking 1 layer
2-06 22 mm punched hole 2 layers
Single
2-08 2 layers
C40/50
2-08 19 mm punched hole 2 layers
2-09 IPE360 beam, 6 m span 19 mm through-welded Single
with 150 mm deep slab C30/37
2-10 1 layer
comprising 19 mm through-welded Pairs
CF80 decking

Figure 3.1 Test setup for beams with CP60 decking, series 2-01 – 2-08 (U Stuttgart)

Figure 3.2 Test setup for beams with CF80 decking series 2-09 – 2-10 (U Luxembourg)

3.2 Description and impact of the research results


All tests were built and performed according to the agreed test configuration. The execution of the
tests mainly followed the instructions given in EN 1994-1-1 B.2.4 for push-out tests:

 25 load cycles up to 40% of the predicted failure load to assess the effects of initial slip (under
load control).
 Increase of loading in steps to the predicted factored load and (in addition to EC4) resting for
each step to determine the influence of short-time relaxation of the concrete, also done at all
push-out tests of WP1.
 Load test to failure under displacement control. The aim was to reach at least a mid-deflection
of span/40.

38
Material properties - Steel
The material properties of the steel beam were provided by ArcelorMittal. The tests provided data
of the yield point and the ultimate tensile strength for upper and lower flange and the web of the
beam-see Table 3.2 .

Table 3.2 Tensile test results (yield point) for IPE sections used in WP 2

Yield point Yield point Ultimate tensile


No. Position strength (MPa)
(MPa) ReH (MPa) ReL

IPE 300 B Bottom flange 403 370 490


IPE 300 H Top flange 411 396 491
IPE 300 C Web 441 418 507
IPE 360 B Bottom flange 382 375 507
IPE 360 H Top flange 381 375 506
IPE 360 C Web 411 399 503

Material properties - Concrete


The material properties of the concrete were provided by the test laboratories of MPA Uni Stuttgart
and Uni Luxembourg in reports that covered the results of concrete compression and tensile tests.
The tensile strength was determined by a 4-point bending test of a concrete prism. Table 3.3 gives
an overview of all tested concrete cubes – and the converted cylinder strengths - belonging to the
series measured at the day of the beam test. All measured values were higher than intended.
Therefore, the tests with nominal C40/50 were tested first to avoid any larger deviation.

Table 3.3 Results of concrete compression and tensile tests

Series 2-01 2-02 2-03 2-04 2-05 2-06 2-07 2-08 2-09 2-10
fc,cube [N/mm²] 56.6 55.4 56.8 55.4 51.8 52.0 80.1 79.2 n.a. n.a.
fc,cyl [N/mm²] 45.2 44.3 45.4 44.3 41.4 41.6 64.2 63.4 48.6* 49.0*
fct,prism [N/mm²] 5.0 4.3 4.2 4.3 4.9 4.5 5.5 5.5 n.a. n.a.
* Tested on cylinders Ø150x300mm, stored in water until testing

3.2.2 Beam Test 2-01


Test 2-01 used diameter 19 mm through-deck welded shear studs placed singly in each deck rib
(see Figure 3.3). The concrete grade was C30/37 and the concrete slab was reinforced only with a
single layer directly positioned above the rib of the steel decking.

Test 2-01 resisted a maximum load Pmax of 388 kN at a corresponding deflection of 53.5 mm
(Figure 3.4). The load-deflection curve of the beam test in Figure 3.5 shows clearly a ductile
behavior of the composite beam. The static failure load was 372 kN, which is only 4% lower than
the maximum load, and means that the effect of relaxation is smaller for the bending resistance
than for the push tests, where it was about 10%. A deflection criterion of span/40 = 125 mm was
exceeded and the test was stopped at a maximum mid-deflection of 145 mm.

Figure 3.3 Construction of CF80 decking series Figure 3.4 Test set-up for CF80 decking
2-09 – 2-10 (U Luxembourg) series 2-09 – 2-10
39
Figure 3.5 Load-deflection curve for beam 2-01

3.2.3 Beam Test 2-02


Test 2-02 used diameter 19 mm shear connectors placed singly in each deck rib through pre-
punched holes. The concrete grade was C30/37 and the concrete slab was reinforced with two
layers. The test set-up is shown in Figure 3.6 and Figure 3.7.

Test 2-02 resisted a maximum load, Pmax of 391 kN at a corresponding deflection of 58.5 mm (see
Figure 3.8). The load-deflection curve of the beam test in Figure 3.7 shows a ductile behavior of
the composite beam. The static failure load was 377 kN, which is only 3.5% lower than the
maximum. The test stopped at a maximum mid-deflection of 140 mm.

Figure 3.6 Test setup of beam 2-02 Figure 3.7 Test 2-02 at failure

Figure 3.8 Load-deflection curve for beam 2-02


40
3.2.4 Beam Test 2-03
Test 2-03 used pairs of diameter 19 mm through-deck welded shear studs per rib. The concrete
grade was C30/37 and the concrete slab was reinforced only with a single layer directly positioned
above the ribs of the steel decking.

The load deflection curve is shown in Figure 3.9. Test 2-03 resisted a maximum load Pmax of
393 kN at a corresponding deflection of 44.5 mm. The load-deflection curve of the beam test in
Figure 3.9 shows clearly a ductile behavior of the composite beam. The static failure load was
375 kN. The test was stopped at a maximum mid-deflection of 140mm.

In a later inspection of the shear connectors, it was found that the weld quality was affected by
welding the shear connectors through the decking close to the ends of the beam flange, which was
not observed in the other tests. When compared to the predicted bending resistance, the failure
load from this test was about 15% lower than expected.

Figure 3.9 Load-deflection curve for beam 2-03

3.2.5 Beam Test 2-04


Test 2-04 used pairs of 19 mm diameter shear studs per rib welded through holes in the pre-
punched decking. The concrete grade was C30/37 and the concrete slab was reinforced only with a
single layer directly positioned above the rib of the steel decking.

Test 2-04 resisted the maximum load Pmax of 454 kN with a corresponding deflection of 49.5 mm
(see Figure 3.10). The static failure load was 435 kN. The test was stopped at a maximum mid-
deflection of 135 mm.

Figure 3.10 Load-deflection curve for beam 2-04

41
3.2.6 Test 2-05
Test 2-05 used single 22 mm diameter shear studs per rib with pre-punched sheeting. The
concrete grade was C30/37 and the concrete slab was reinforced with a single layer positioned
above the ribs of the decking. Beam 2-05 resulted in a maximum load Pmax of 406 kN at a
deflection of 54mm (see Figure 3.11). The static failure load was 387 kN. The test was stopped at
a maximum deflection of 140 mm.

Figure 3.11 Load-deflection curve for beam 2-05

3.2.7 Beam Test 2-06


Beam 2-06 used single 22 mm diameter shear studs per rib with pre-punched sheeting. The
concrete was C30/37 and the concrete slab was reinforced with two layers. Test 2-06 resulted in
maximum load Pmax of 418 kN at a deflection of 60 mm (see Figure 3.12). The static failure load
was 402 kN. The test was stopped at a maximum deflection of 140 mm.

Figure 3.12 Load-deflection curve for beam 2-06

3.2.8 Beam Test 2-07


Beam 2-07 used single 22 mm diameter shear studs per rib welded through holes in the pre-
punched decking. The concrete grade was C40/50 and the concrete slab was reinforced with two
layers. Test 2-07 resulted in maximum load Pmax of 430 kN at a deflection of 51 mm (see Figure
3.13). The load-deflection curve of the beam test shows clearly a ductile behavior of the composite
beam. The static failure load was 412 kN. The test was stopped at a maximum deflection of
165 mm.

42
Figure 3.13 Load-deflection curve for beam 2-07

3.2.9 Beam Test 2-08


Test 2-08 used single studs per rib of 19 mm diameter with pre-punched sheeting. The concrete
grade was C40/50 and the concrete slab was reinforced with two layers. Test 2-08 had a maximum
load Pmax of 408 kN at a corresponding deflection of 52 mm (see Figure 3.14). The load-deflection
curve of the beam test shows clearly a ductile behavior of the composite beam. The static failure
load was 397 kN. The test was stopped at a maximum deflection of 142 mm.

Figure 3.14 Load-deflection curve for beam 2-08

3.2.10 Beam Test 2-09


Test 2-09 used single shear studs per rib of 19 mm diameter which were welded through the steel
decking. The test set-up is shown in Figure 3.15 and Figure 3.16. The concrete grade was C30/37
and the slab was reinforced with a single layer. Test 2-09 resulted in a maximum load Pmax of
219 kN at a corresponding deflection of 54.8 mm (see Figure 3.15). The load-deflection curve (see
Figure 3.17) shows a ductile behavior of the composite beam. The static failure load was 213 kN.
The test was stopped at a maximum deflection of 180 mm.

43
Figure 3.15 Test set-up of specimen 2-09 Figure 3.16 Test 2-07 at maximum deflection
250

200
Point load [kN]

150

100

50

0
0 20 40 60 80 100 120 140 160 180 200
Mid-span deflection [mm]

Figure 3.17 Load-deflection curve for beam 2-09

3.2.11 Beam Test 2-10


Beam 2-10 used pairs of shear studs per rib of 19 mm diameter. The studs were welded through
the decking. The concrete grade was C30/37 and the slab was reinforced with a single layer.
Test 2-10 resulted in maximum load Pmax of 240 kN at a deflection of 68.8 mm (see Figure 3.18).
The load-deflection curve shows a sudden increase of the deflection at 207 kN. At this load rib pry-
out failure was observed. For further loading, the test showed a ductile behavior. The static failure
load was 233 kN. The test was stopped at a maximum deflection of 164 mm.

Load-deflection curve for test 2-10


300

250
Point load [kN]

200

150

100

50

0
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170
Mid-span deflection [mm]

Figure 3.18 Load-deflection curve for beam 2-10


44
3.3 Development of end slip
The development of slip was measured at the ends of the beam as well as throughout the span of
the beams, as shown in Figure 3.19. Figure 3.20 presents the slip results in test 2-01. In general,
the development of slip was comparable for all beam tests. At first, the slip developed
proportionate to the increase of load. At about 40% of the maximum load, slip began to increase
disproportionately. Although the beams were manufactured symmetrically and installed centrically
below the testing rig, one side of the beam (for example the right hand side of beam test 2-01)
failed first and developed more slip than the other side.

Figure 3.19 Measurement of slip along the beam

Figure 3.20 Slip development for beam test 2-01

Figure 3.22 shows the test beam 2.01 which was cut along the edges of the flange to show the
failure mode of the concrete in the ribs of the CP60 decking. The local view of the deformed 19 mm
diameter shear connector that was welded through the decking is shown. The higher end slip on
one side of the beam was possibly because some studs on one side are weaker (possibly due to
inferior welding quality and the variability of concrete).

The relevant positions for yielding in the lower flange of the steel beam are shown in Figure 3.22
also for beam test 2-01. Yielding of the steel beam occurred first in the mid-span region of the
beam and then below the load introduction points. Because 6 of 10 tests failed at left hand side
(A) and 4 of 10 at right hand side (B), it can be concluded that there was no systematic error
during manufacturing of the specimens and/or execution of the beam tests.

Figure 3.21 Results of cutting specimen 2-01 on both ends of the beam

45
Load-strain curve 2-01
Development of yielding in the lower flange

450

400

350

300
Load [kN]

250
DMS 31

200 DMS 41
DMS 51

150

100 P P

50
DMS 31 DMS 41 DMS 51

0
0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000
Strain [µm/m]

Figure 3.22 Strain in the lower flange of the steel beam for beam test 2.01

3.4 Comparison of beam tests with plastic bending resistance for partial shear
connection
Table 3.4 presents the shear connector resistances that are input into the plastic analysis of the
composite beam together with the other material properties described earlier. Six methods are
presented in this table;
 The test resistance of the shear connector PR obtained from WP1
 The characteristic resistance, which for single tests is taken as 0.9 PR
 The characteristic resistance calculated to Eurocode 4 using the kt value for the particular deck
profile obtained from clause 6.6.4.2 of EC4 using the nominal height of the shear connector
and the shear connector strength, fu of 450 N/mm2
 The characteristic resistance calculated using the kt value for the deck profile using the as-
welded height of the shear connector (taken as height, hsc -5mm)
 The characteristic resistance obtained from the method in the thesis of Konrad (Uni Stuttgart).
 The characteristic resistance obtained from the method in the thesis of Nellinger (Uni
Luxembourg).
As can be seen from the data in this table, there is a considerable range in the characteristic
resistances among all of the methods and the Eurocode 4 value is often significantly higher than
the test resistances.

46
Table 3.4 Shear connector resistances Pstud [kN] as input in the plastic analysis of the beam
tests (ST)

Shear connector ST ST ST ST ST ST ST ST ST ST
resistance from: 2-01 2-02 2-03 2-04 2-05 2-06 2-07 2-08 2-09 2-10

WP 1 Test failure,PR 62.4 51.6 61.5 54.7 86.7 83.1 101.5 52.5 67.9 40.8
WP 1 Pk = 0.9 PR 56.2 46.4 55.4 49.3 78.0 74.8 91.4 47.3 61.1 36.7
EC4 characteristic, Pk 72.7 72.7 51.4 97.5 97.5 97.5 97.5 72.7 69.1 48.8
EC4 characteristic, Pk 69.2 69.2 49.0 49.0 92.8 92.8 92.8 69.2 61.4 43.4
(using hsc-5mm)
Konrad, PKonrad 58.0 46.6 37.7 37.3 60.5 60.6 62.5 47.2 73.4 58.6

Nellinger- PN-WP5.1 49.2 48.9 27.0 26.8 53.7 53.7 59.3 59.3 51.9 29.1

Table 3.5 presents the degree of shear connection in each test based on the characteristic shear
connector resistance, 0.9 PR from WP1 and the calculated plastic bending resistance MR,calc obtained
from partial shear connection principles using the number of shear connectors from the support to
the load points. The model factor is the ratio of the bending resistance in the test to the bending
resistance obtained from this calculation. A model factor of more than 1.0 indicates that the plastic
method is accurate. Because the plastic model is an idealisation based on fully developed plastic
stress blocks, it would be expected that model factors slightly less than 1.0 would be obtained for
these relatively short span beams subject to point loads rather than uniform loading.

Table 3.5 Model factor (MTest/MR,calc) for the beam tests in WP2

Test Degree of WP 1 Push test EC4


shear results Characteristic, Pk Nellinger
connection Konrad
Shear connector Method
(Pk = 0.9PR) P = P method
k R height WP 5.1
Pk = 0.9PR
=hsc hsc-5mm
ST 2-01 21% 0.96 0.99 0.95 0.96 0.98 0.94
ST 2-02 17% 1.02 1.04 0.96 0.97 1.04 0.95
ST 2-03 42% 0.83 0.85 0.88 0.89 0.92 0.92
ST 2-04 37% 0.98 1.00 1.02 1.03 1.07 1.10
ST 2-05 29% 0.92 0.95 0.92 0.93 1.01 0.96
ST 2-06 28% 0.96 0.99 0.96 0.97 1.04 0.99
ST 2-07 34% 0.94 0.96 0.98 0.98 1.05 0.98
ST 2-08 18% 1.06 1.09 1.00 1.02 1.09 0.95
ST 2-09 15% 0.91 0.93 0.92 0.95 1.07 0.95
ST 2-10 17% 0.95 0.97 0.94 0.96 0.97 1.03
Mean value (all tests) 0.95 0.98 0.95 0.97 1.02 0.98

Mean value (ex. ST2-03) 0.97 0.99 0.96 0.98 1.04 0.98
Standard deviation 0.045 0.046 0.032 0.029 0.038 0.051
Coefficient of variation 0.047 0.047 0.033 0.030 0.037 0.049

The calculated bending resistances used the shear resistances presented in Table 3.4 as input.
Because test 2-03 failed because of poor weld quality of the pairs of shear connectors, it is either
considered or eliminated in calculating the mean of the test results.

The model factor results show that when the characteristic value of the shear connector resistance
is used based on the push tests, the model factor is 0.99 on average (ignoring test 2-03) with a
coefficient of variation of 0.046. As the plastic analysis model is an idealisation of the true
behaviour of the beam, a model factor of 0.95 is considered to be acceptable in the calibration of
the design rules for beams with low degrees of shear connection.

47
The model factor based on the characteristic resistance to Eurocode 4 is 0.96 on average (ignoring
test 2-03) with a coefficient of variation of 0.033. This implies that the Eurocode 4 shear connector
resistances are unconservative by about 10% on average. The model factors according to the
Konrad method is 1.04 on average (ignoring test 2-03) and the model factors according to the
Nellinger method is 0.98 on average.

These results are summarised in Figure 3.3(a) (without test 2-03). The degree of shear connection
varied between 15% and 43%. As seen in Figure 3.23(b) for all tests, there was little correlation
between the model factors and degree of shear connection in terms of the plastic bending
resistance model. Indeed, tests with lower degrees of shear connection often had higher model
factors.

As an additional analysis, the shear connector resistance to Eurocode 4 was calculated for the as-
welded height of the shear connector (taken as hsc – 5 mm). This increased the model factor to
0.98 (without test 2.03) and the coefficient of variation reduced to 0.03. This may be one way of
reducing any un-conservatism in the Eurocode 4 shear connector resistances.

650 0,7
WP 1 Tests 90%
625 Mtest / MR,calc
KONRAD Prk
600 0,6 EC4 Prk
575 Mtest / 0,95*MR,calc NELLINGER Prk

Degree of shear connection [-]


550 0,5
525
Mtest [kNm]

500 0,4
475
450 0,3
425
400 0,2

375 WP 1 Tests 90%


KONRAD Prk 0,1
350
EC4 Prk
325 NELLINGER Prk
300 0,0
300 325 350 375 400 425 450 475 500 525 550 575 600 625 650 0,80 0,85 0,90 0,95 1,00 1,05 1,10
MR,calc [kNm] Mtest/Mcalc [-]

(a) Test moment versus calculated bending (b) Degree of shear connection versus model
resistance - excluding test 2-03 factor
Figure 3.23 Results of the plastic resistance of the beam tests using the characteristic shear
connector resistance (Pk= 0.9 Ptest or Pk to EC4)

3.5 Concluding remarks for WP2


The 10 beam tests showed that the plastic bending resistances of the composite beams could be
developed when using the characteristic shear connector resistances obtained from the push tests
(this is taken as 0.9 x static test result for single tests). One beam test (2-03) failed prematurely
because of poor weld quality when pairs of shear connectors were used. If this test result is
eliminated, the variation in the results was small relative to the prediction based on plastic bending
resistance for the particular degree of shear connection. However, test 2-03 has highlighted the
importance of good control of welding for pairs of shear connectors when placed close to the ends
of the flange.

The comparison using the characteristic resistance of the shear connectors to Eurocode 4 is not as
good, but still leads to a calculated bending resistance that does not exceed the test bending
resistance by more than 5% on average despite the low degrees of shear connection in the tests.
The coefficient of variation was only 3% in the comparison with Eurocode 4.

3.5.1 Exploitation and impact of the research results


This series of beam tests gives a strong indication of the importance of the shear connector
resistance obtained from push tests and shows that the partial shear connection method may be
used safely despite the low degree of shear connection in the tests. The results will be used to
make improvements in Eurocode 4.

The low failure load recorded for test 2-03 highlights the importance on the control of the welding
procedure for pairs of shear connectors welded through the decking.

48
4 WP3: Tests on long span asymmetric composite beams

4.1 Objectives of WP3


In this WP, a series of long span beam tests was devised to investigate the performance of
composite beams of asymmetric cross-section and with low degrees of shear connection which
complimented the shorter span tests in WP2. The details of the tests were as follows:

 15.3 m span cellular beam of 650 mm depth with high asymmetry in flange areas.
 Asymmetric loading test on the same cellular beam to determine the shear transfer at the
1.05 m long elongated opening.
 11.2 m span fabricated section of 450 mm depth with asymmetry in bottom to top flange areas
of 1.5 to 1.
The beams were cast with projecting out-riggers to support the decking so that the self-weight
loads were supported by the steel beam. When cast, the beams were tested under 8 point loads to
simulate uniform loading. Shear connectors were welded at 300 mm spacing in every rib of the
80 mm deep deck profile and the degree of shear connection was 36% in the cellular beam test
and 33% in the fabricated beam test, both of which were well below the minimum required to
Eurocode 4 for long span beams.

The main objectives of these long span beam tests were to investigate:

 The ability of the composite beam to develop its plastic bending resistance with low degrees of
shear connection (less than the required limit given in the Eurocode 4).
 The effect of unpropped construction on the longitudinal shear connector forces.
 The difference in response between the solid-web beam and cellular beam.
 Local composite action at the elongated openings ta mid-span.
 The effect of partial shear connection and the regular openings on the additional deflection of
composite beams.

4.2 Comparison of initially planned activities and work accomplished


The beam types that were chosen for the long span beam tests were designed carefully to
demonstrate a flexural mode of failure rather than shear and to reflect the practical nature of
modern composite construction. These tests were chosen as follows:

 The first test was on a 15.3 m span composite asymmetric cellular beam manufactured by
cutting and re-welding an IPE 450 top chord and an HEB 360 bottom chord to create a beam
with multiple circular openings. The asymmetry of bottom to top flange areas was 2.43:1, which
is close to the extreme of the application range.
 The second test was on the same cellular beam but in this case loading was applied
asymmetrically to cause failure by Vierendeel bending at an elongated opening when subject
also to a low degree of shear connection.
 The third test was on a 11.2 m span solid web composite beam fabricated from three plates and
with a ratio of bottom to top flange areas of 1.5:1, which is in the intermediate span and
asymmetry range.
The other parameters in the tests were:

 150 mm deep composite slab.


 80 mm deep composite decking (CF80 x 0.9 mm thick steel as tested in WP1).
 Slab width of 3m (< span/4 = 3.8 m to conform to the lab space) for the 15.3m span and 2.8m
(=span/4) for the 11.2 m span.
 Loading applied at 8 points to simulate uniform loading.
 Use of S355 steel in the beams (measured strengths were obtained).
 Use of steel outriggers to support the decking so that self-weight loads were applied to the steel
beam to simulate un-propped construction.
4.3 Description and impact of the research results

Cellular beam test


The cellular beam spanned 15.3 m between the supports and comprised two steel sections. The
top T-section was 305 mm deep and was cut from an IPE 450 section, while the bottom T-section
was 260 mm deep and was cut from a HEB 360 section. This gave a total section depth of 565 mm

49
and a length-to-depth ratio of 27. Hence the beam was highly asymmetric with a bottom flange
area to top flange area ratio of 2.43:1. The diameter of the cell openings was 425 mm with a
centre-to-centre spacing of 680 mm. The cross-section dimensions are shown in Figure 4.1.

Figure 4.1 Cross-section dimensions of the asymmetric cellular beam

The elongated opening at the mid-span was 1105 mm long and was made by removing the central
web-post. Some openings at or near to the load positions were fully or half in-filled to avoid local
failure and the arrangement of the openings was symmetrical about the mid-span.

The steel decking (CF80) is illustrated in Figure 2.6(b). It was 80 mm deep and 0.9 mm thick with
a 15 mm re-entrant stiffener on top. Single 19 mm diameter x 125 mm long shear connectors were
placed at 300mm spacing in each rib, making 51 in total. The critical cross-section was at the load
point at 7/16 of the span and 22 shear connectors were placed from the support to this point. The
welded height of the shear connectors was 120 mm. A193 mesh reinforcement (7 mm wire;
200 mm  200 mm) was placed at the same level as the top of the welded shear connector.

The shear connector resistance was taken as 70 kN based on the tests for single shear connectors
in WP1. The degree of shear connection was 36% at the critical cross-section compared to a
Eurocode 4 requirement of 88% for the degree of asymmetry and the steel strength. At this low
degree of shear connection, it was expected that end slips would be very high at the plastic
bending resistance of the cross-section at an opening. The results were expected to confirm that
cellular beams can be designed for a lower degree of shear connection than the equivalent solid
web beams because it is not necessary to develop plasticity in the ‘missing web’ of the beam. This
in investigated further in WP5.3.

The concrete slab width was chosen as 3.0 m, primarily for practical reasons and because this is
less than the theoretical effective slab width for this beam span. Figure 4.2(a) shows the steel
cellular beam prior to placing the decking and Figure 4.2(b) shows the view of the outriggers used
to support the ends of the decking. Figure 4.3 shows the loading system.

(a) View of cellular beam (b)Out-riggers connected to beam


Figure 4.2 Set-up for uniformly distributed load test on cellular beam

50
1907.5mm 160 mm
1600
3815 mm 3815 mm 1907.5mm

15260 mm
15660 mm
Figure 4
4.3 Set-up for uniformly distributed load test on cellular beam

4.3.2 Test results for cellular beam


The load-deflection results for the 5 cycles of loading are presented in Figure 4.4. Results were
expressed a an equivalent uniform load by dividing by the slab area of 45.8m 2). The deflections
were cumulative as there was a noticeable initial deflection after the second load cycle to
7.5 kN/m2. The failure load was 17.2 kN/m2 in addition to the self-weight of the slab, test beam
and loading beams which were calculated to be 3.6 kN/m2. This is equivalent to an applied moment
of 1816 kNm.

The end slips between steel beam and concrete slab at both ends of the composite beam are
presented in Figure 4.5. The slips were 13.5 mm and 8.5 mm at each end under the maximum
applied load of 17.2 kN/m2. Non-linearity in the load-slip curve occurred for loads exceeding
8 kN/m2 or approximately half of the eventual failure load.

The strain in the bottom flange is shown in Figure 4.5. The strain at failure was only about 20%
higher than the yield strain.

Figure 4.4 Applied load versus mid-span deflection for cellular beam test

51
Figure 4.5 Applied load versus end-slip measured by LVDT_11 for cellular beam test

Figure 4.6 Applied load versus strain in bottom flange at mid-span for cellular beam test

The calculated plastic bending resistance of the cross-section at an opening was calculated to be
1798 kNm for the calculated 36% degree of shear connection. This corresponds to 99% of the test
bending resistance and is considered to be acceptable for the low degree of shear connection in the
test.

4.3.3 Test on cellular beam with asymmetric loading


In this series of tests, loading was applied off-centre to take account of asymmetric loading and
this caused shear in the elongated opening, as shown in Figure 4.7. This was known as a ’shear
test’ and had the objective of investigating the behaviour of the shear connectors in shear and
tension around the elongated openings. Figure 4.8 shows the relationships between load and
deflection at the loading positions and mid-span.

The failure load (maximum applied load) was 405 kN, which included the weight of the spreader
beams but not the self-weight of the concrete slab. Non-linearity in the load-deflection curve
occurred at loads over 300 kN. The deflected shape of the opening is shown in Figure 4.9. The
differential deflection across the opening is shown in Figure 4.10.

52
Figure 4.7 Testing arrangement for asymmetric load test

Figure 4.8 Applied load versus deflection at loading position for the asymmetric load test

Figure 4.9 Deflected shape of the beam and elongated opening for the asymmetric load test

53
Figure 4.10 Load versus deflection difference across the opening from LVDT_4 to LVDT_6 in the
shear test

4.3.4 Measured strains in cellular beam test


As discussed previously, the strains in the steel beam were measured during the casting of the
concrete. For the cellular beam test, the strain in the top flange at the mid-span was 497µ and
the strain in the bottom flange was 260µ. If these strains are added to the total accumulated
strains during the uniform load test presented in Figure 4.11, the strains in the top and bottom
flanges increased from 1205µ and 2295µ to 1702µ and 2555µ respectively. The position of the
neutral axis can be estimated using values of strain in the top and bottom flanges of the beam.

1035με 1035με
150 mm 106 με 106 με
497με
1205με 1702με

195mm 226 mm
371 mm
565 mm

260με 2295με 2555με


After concreting Measured during test Combined strains
Figure 4.11 Strain distribution through cross-section depth corresponding to measured strains for
cellular beam test 1

The depth of the neutral axis is 371 mm from the top of the beam after the pouring of the wet
concrete. Using the maximum accumulated strains of 1205μ and 2295μ, the neutral axis depth
during the test was estimated to be 195 mm. However, when the strains from the weight of the
concrete were also included, the position of the neutral axis reduced to 226 mm below the top of
the beam.

4.3.5 Details of asymmetric fabricated beam test


A 450 mm deep asymmetric beam of 11.2 m span with a ratio of flange areas of 1.5 was tested as
being typical of fabricated beams. The section comprised a 10mm thick top flange and web and
54
15 mm thick bottom flange, and both flanges were 180 mm wide. The slab was 2.8 m wide (=
span/4). The beam was subject to 8 point loads to simulate uniform loading that was progressively
increased in cycles of loading to determine the effects of slip in the serviceability and ultimate load
ranges.

The slab was 150 mm deep and used the CF 80 decking as in the other long span tests. Shear
connectors (125 mm height and 19 mm diameter) were welded through the decking at a spacing of
one per deck rib at 300mm spacing. The characteristic resistance of the shear connectors was
obtained as 75 kN from push tests on the same 80mm deep deck profile (for C30 concrete), and
this was down–rated to 68 kN for the measured C25 concrete in the beam test. A total of 16 shear
connectors was placed from the support to the critical cross-section at the load point at 7/16 of the
span.

The same system of construction was used to ensure that the loads from the concrete slab were
supported by the steel beam as shown for the cellular beam. The out-riggers were removed when
the concrete had gained its full strength. The stress in the bottom flange due to the self-weight of
the slab and beam was measured at about 29% of the steel yield strength. The beam under load is
shown in Figure 4.12.

Figure 4.12 Arrangement of actuators and load-spreader beams for the 11.2m span fabricated
beam test

4.3.6 Test results for asymmetric beam


Three cycles of increasing load were applied up 50% more than normal service loading and a
further 4 cycles were applied to failure. The maximum displacement at the end of the test was
220 mm (=span 50). The beam after un-loading from the failure load is shown in Figure 4.13, in
which the residual deflection was about 120 mm.

The full load-deflection curve is shown in Figure 4.14, which indicated that full plasticity had
developed at failure. The applied loading at failure was 18.0 kN/m2 plus 3.6 kN/m2 for the self-
weight of the slab and beam and spreader beams. This was based on a loaded area of 31.3m2. This
was equivalent to a bending moment of 948 kNm. The predicted failure moment was 965 kNm,
based on measured material strengths, which is only 2% higher and within the normal margin of
acceptance for composite beam tests.

The load-deflection curve for the fifth load cycle shows linear behaviour up to a load of 10 kN/m 2,
which is twice normal service loading. A permanent deflection of about 7 mm was measured after

55
unloading from this cycle, which indicates that irreversible deformation of the shear connectors had
occurred.

The load-slip behaviour of the shear connectors is shown in Figure 4.15. Up to service load levels,
the slip in the shear connectors was less than 0.5mm, and it was observed that slip increased more
rapidly at a load of about 7 kN/m2. At a load of 12 kN/m2, the end slip was about 3 mm. The
limiting slip of 6 mm for ‘ductile’ shear connectors was passed at a load of 15 kN/m 2, and from this
point, slip increased rapidly to a maximum of 19 mm at the end of the test, which demonstrated
high deformation capacity.

Figure 4.13 Deflection of composite fabricated beam after unloading from failure

Figure 4.14 Relationship between load and mid-span deflection for fabricated beam

Figure 4.15 Relationship between load and end-slip for fabricated beam

56
4.3.7 Measured strains in asymmetric beam test
For the 11.2 m fabricated beam test, significantly higher strains were observed at mid-span in the
flanges on the right hand side of the steel beam than on the left hand side. The strains measured
during all loading cycles are shown in Figure 4.16 and Figure 4.17 for the left hand and right hand
sides of the beam respectively. The strain distribution through the cross-section is shown in Figure
4.18, and the neutral axis position was 144 mm from the top flange.

Figure 4.16 Relationship between load and strain in bottom flange for the fabricated beam test

Figure 4.17 Strain distribution through cross-section corresponding to measured strains and average
measured flange strains recorded for the fabricated beam

4.4 Comparison of deflections of the tests


For the fabricated beam test, at a typical imposed load of 5 kN/m2, the measured deflection of the
beam was 17 mm (eliminating the residual deflection after the earlier load cycles). The calculation
of the effective inertia and end slip is given below using the following data:

e = Es/Ec = 6 (for short-term loads)


As = 8.75 x 103 mm2
Ac = 70 x 2800 = 196 x 103 mm2
Is = 263  106 mm4
Ic = 80  106 mm4 (for hc = 70mm)
ys = 237 mm for steel beam (measured from top flange)

57
The composite second moment of area with rigid shear connectors is Icomp = 1104 x106 mm4. The
shear connector stiffness per unit length is k/ssc = 70/300 = 0.23 kN/mm2. The effective second
moment of area of the composite beam with flexible shear connectors is obtained from
equation (9.1) in WP6 as:

 3 
237  35  802  196 x10 

6
80 x10  6 = 821 x 106 mm4

I  236x106  
eff 6  3 2  3 
1  196x10   π  210  196x10 
 6 x8750  11200 0.23  6 
  

The effective inertia Ieff is therefore 74% of the composite inertia, Icomp. For an imposed load of
5 kN/m2, the applied bending moment is 220 kNm and the end slip is obtained from equation (9.2)
in WP6 as 0.53 mm. It follows that the maximum force in the shear connectors is: 0.53 x 70 = 37
kN at this load, and so the shear connectors were still elastic. Using this stiffness, the shear
connectors would reach their elastic limit at a load of about 10 kN/m 2. This agrees well with the
fabricated beam test in WP3. Their elastic limit is reached at 46% of the failure load of the beam.

The calculated deflection of the fully composite beam with rigid shear connectors is 12 mm and the
deflection of the composite beam with flexible shear connectors (using k= 70 kN/mm) is 16.2 mm.
Therefore, the additional deflection due to end slip is 35% of the deflection of the composite beam
with rigid shear connectors. The test deflection was approximately 16 mm at this load, which is in
good agreement with the theoretical deflection using this flexibility, as seen in Figure 4.18.

Figure 4.18 Load-deflection cycle for the asymmetric fabricated beam up to twice working load
and comparison with the theory

The test deflections for the two beams at an imposed load of 5 kN/m2 are compared in Table 4.1
with the theory presented in this paper and with the approximate formulae in BS 5950-3 and the
American AISC codes. The stiffness of a cellular beam is based on the proportionate length of the
openings. No formulae for additional deflections due to slip are given in Eurocode 4 and so the
comparison can be made only with the deflection due to the fully composite stiffness. The effective
stiffness is calculated for two shear connector stiffness of k= 70 and 100 kN/mm.

It is shown that the Code methods are conservative for the low degrees of shear connection in the
tests. The BS 5950-3 method over-predicts the test deflections by 12 to 21%. The AISC code
formula over-predicts deflections by 8 to 12%.

58
Table 4.1 Comparison of the test results on long span beams with Code methods

Test beam Degree of Measured Deflection of Eqn (17) using Existing Code
shear deflection composite shear connector Methods
connection in test at beam stiffness of:
in test 5 kN/m 2 without end
k = 70 k = 100 BS AISC
slip
kN/mm kN/mm 5950-3 Code
11.2m span 33% 16.0mm 12.0mm 16.2mm 16.0mm 18.0mm 17.2mm
fabricated
beam
15.3m span 38% 22.0mm 17.3mm 23.4mm 22.3mm 26.6mm 24.7mm
cellular beam

4.5 Concluding remarks for WP3


A series of full-scale beam tests was conducted to evaluate the response of two un-propped
composite beams with low degrees of shear connection. They were tested under uniformly
distributed loading and the cellular beam was also tested under asymmetric loading. The findings
are summarised as follows:

 Under simulated uniformly distributed loading, the cellular beam resisted a maximum load of
3.4 × design working load prior to failure with a degree of shear connection of 36% below the
88% minimum to Eurocode 4 Clause 6.6.1.2. This suggests that the rules for the minimum
degree of shear connection could be relaxed. The fabricated beam resisted a similar level of
loading prior to failure.
 At the maximum uniformly distributed load, the bottom flange of the cellular beam had reached
yield, but the strains in the concrete were considerably lower than the failure strain. The
bending resistance of the beam exceeded the design prediction to EC4 despite its low degree of
shear connection.
 No concrete crushing was observed in the 11.2 m fabricated beam test, and only a longitudinal
crack occurred at the level of the shear connectors along the beam length. The tensile strains
measured in the bottom flange were several times greater than the yield strain for a
considerable portion of the beam length. Large compressive strains exceeding the yield strain
were also observed in the top flange of the beam.
 Ductile behaviour of the shear connectors was confirmed in the cellular beam test by the
observed end slips of 8.5 mm and 13.5 mm under uniformly distributed loading, which
exceeded the minimum of 6 mm specified in Eurocode 4 Clause 6.6.1.1. An end slip of
19.2 mm was measured in the fabricated beam test.
 A mid-span vertical deflection of span/100 was a suitable criterion for failure for the cellular
beam, as evidenced by the reducing load and excessive deflections that occurred in the
subsequent loading cycle. This corresponded to a maximum equivalent uniform load of
17.2 kN/m2. The load-deflection behaviour was linear up to 7.5 kN/m2, and little residual
deflection was observed after unloading at this stage of the test.
 In the asymmetric load test, the beam supported a maximum point load of 405 kN, which was
45% higher than the predicted resistance of the steel cross-section. Neglecting the reduction
factor applied to account for pull-out forces, the bending maximum load would still exceed the
predicted load by 28%. This suggests the need for modifications to the predicted Vierendeel
bending resistance for composite cellular beams.
The findings suggest that Eurocode 4 design recommendations could be modified to better exploit
the properties of un-propped composite beams, particularly with regard to the required degree of
shear connection. Numerical analysis using the finite element software package ABAQUS in WP5.2
supported the recommendations following from this research project.

59
60
5 WP4: Composite floor plate test to investigate interactive effects

5.1 Objectives of WP4


The objective of this composite floor plate test was to investigate how the in-plane or membrane
effects of the composite floor slab beam system that reduces the tendency of the longitudinal
splitting of the concrete slab along the line of the primary beams and therefore increases the load
capacity of the structural system. This is important in cases where the steel decking is
discontinuous when it is orientated parallel to the beams. In this study, the amount of transverse
reinforcement to transfer local forces from the shear connectors is compared to the requirements
of EN 1994-1-1.

The research investigated the in-plane forces developed in the slab due to the restraining action of
the floor plate that is held in position by the peripheral composite beams. The tendency for
cracking along the primary beam and at the peripheral beams was measured. To be realistic, the
floor plate test consisted of a grillage of primary beams, secondary beams, and columns. The
precise details of the test were designed to investigate the potential for longitudinal splitting of the
concrete slab along the primary beam and membrane effect of the concrete slab. The test was
devised so that the internal primary beam would fail before the secondary beams.

5.1.1 Comparison of initially planned activities and work accomplished


The floor plate test was slightly smaller than given in the proposal in order for it to conform to the
space available in the laboratory and to the constraints of the reaction points on the lab floor. The
plate test was designed so that the longitudinal force transferred by the primary beams was
relatively high by designing the beam for full shear connection, and the transverse reinforcement
was taken as the minimum of 0.2% of the concrete cross-sectional area over the decking.

The following additional tests were performed:

 The proposal envisaged that a test would be performed on one edge beam. However, it was
decided that it would be useful scientifically to test the two similar edge beams with and without
U bars around the shear connectors in order to evaluate the difference in load capacity and to
determine the shear connector resistances by back-analysis of the test failure loads.

 The internal columns were fabricated 200 mm short of the foundation in order to allow for a
load cell to be positioned below. After completion of the test series, the load cell was removed
and the floor was loaded without support to the internal column. This ‘pseudo-robustness’ test
on the damaged floor plate was continued until a deflection of 200 mm was reached.

5.2 Description and impact of the research results

5.2.1 Description of the tests


The composite floor plate dimensions were chosen as 10.6 m x 4 m. The composite floor plate
consisted of 9 beams (3 primary beams and 6 secondary beams) and 6 columns, as shown in
Figure 5.1. The central primary beam was IPE 270 in S355 steel and spanned 3.6 m between the
flanges of the two central columns. The other two edge primary beams had the same dimensions
and material properties as the central primary beam.

The beams were manufactured with end plate connections where they were connected to the
columns and where the secondary beams were connected to the primary beams. The steel decking
was chosen as CP60 with ribs at 207 mm spacing (CP 60 was tested extensively in WP1). It was
placed continuously over the secondary beams but was discontinuous over the primary beams. The
shear connectors were welded to the primary beams at a spacing of 200 mm so that full shear
connection was developed in order to maximise the longitudinal shear force that was transferred
and therefore to maximise the splitting effect that could occur.

The mesh reinforcement was placed in the floor slab at a depth just below the shear connector
head. The edge distance of the shear connectors was 50 mm from the slab edge and on one edge
beam U bars were placed around the shear connectors.

The test to fail the primary beam was carried out with 2 point loads applied to each of the
secondary beams so that the primary beams were loaded via the secondary beams, as would be
the case in practice. This was chosen as an alternative to loading directly through the slab in order
to minimise the load-sharing effects among the beams due to the two-way spanning action of the
slab.
61
Figure 5.1 Beam arrangement and floor plate dimensions

The two internal secondary beams were IPE 300 in S355 steel and were connected to the webs of
the supporting beams (one end to the central primary beam and the other end to the edge primary
beam). All of the edge beams (including edge primary beams and edge secondary beams) were
IPE 270 sections. The central primary beam (IPE 270) was designed to be 10% weaker than the
internal secondary beams (IPE 300) so that failure might occur first in the central primary beam.

5.2.2 Test series for floor plates


The floor plate was designed so that the longitudinal force transferred by the central primary beam
was relatively high, and the transverse reinforcement provided was the minimum practically
possible, which was A142 mesh (0.2% of the slab cross-sectional area). From the design
calculations, the central primary beam had a 35% degree of shear connection and the internal
secondary beams had 26% degree of shear connection. Both were less than the minimum of 40%
required to Eurocode 4. At one longitudinal edge beam, U bars were used to provide local
anchorage, whereas at the other edge, no additional reinforcement other than the mesh was used.

The floor beam and column arrangement is shown in Figure 5.2. The IPE 300 internal secondary
beams were 5.24 m long between the webs of the supporting primary beams and the IPE 270
primary beams were 3.6 m long between the flange faces of the columns. The columns were
62
approximately 1 m high and the two central columns were detailed to be shorter than the other
4 corner columns so that load cells could be placed underneath them to monitor the loads. The two
central columns were tied to prevent outward movement since it was possible that a small moment
might be generated in the columns through the connections.

Figure 5.2 Layout of beams in floor plate test

5.2.3 Test series on floor plate


The tests were carried out in four stages:

 The central primary beam was first loaded through the internal secondary beams by applying
two point loads per secondary beam.
 Then the two point loads were moved to the edge secondary beams without U bars in the slab.
 After testing the edge secondary beams without U bars, the two point loads were moved to the
edge secondary beams with U bars to obtain the difference in the failure load due to the U bars.
 Finally a robustness test was performed on the side of the edge secondary beams without
U bars in which the support to an internal column was removed.
In the first test, the loading of the central primary beam at the mid-span was through the two
internal secondary beams as shown in Figure 5.3. The second test had the objective of identifying
whether longitudinal cracking had occurred in the slab at the edge beams, since it is normally
required to provide additional reinforcement in the form of U-shaped bars to control splitting of the
concrete locally. For the third test, the loads were applied to the edge secondary beams with
additional U bars to compare with the case without additional U-bar reinforcement. The
reinforcement and decking details are shown in Figure 5.4 and Figure 5.5.

After completing the above test series, a pseudo-robustness test was performed by removing the
support under a central edge column. This test allowed a column deflection of just less than
200mm which was not sufficient to cause failure of the connections, but showed that the already
damaged floor slab could resist significant load compatible with a robustness scenario as envisaged
by Regulations.

63
Figure 5.3 View of load test on secondary beam

Figure 5.4 Details for shear connectors at primary beam and edge beam

64
Figure 5.5 Details of U-bar arrangement at edge beam

5.3 Floor plate results


The deflection measurement positions are shown in Figure 5.6. Absolute deflections are plotted but
the net deflection of the secondary beams was obtained by deducting the average deflection of the
primary beams.

Figure 5.6 Deflection (linear voltage displacement transducers) positions LVDT 1 to 16

5.3.1 Loading of primary beam through internal secondary beam


The loading system through the secondary beams is shown in Figure 5.7. The deflected shape of
the secondary beams at failure is shown in Figure 5.8. Based on the deflections recorded for the
edge beams, it was estimated that the load resisted by the secondary beams was 80% of the
applied load and this load was transmitted to the primary beams. The remaining load was
transferred via the edge beams to the columns due to flexure of the floor slab. This is the basis of
the comparison between tests and theory, which was confirmed in WP4.3 by a finite element model
of the whole floor plate.

For the edge beam tests, it was estimated that the load resisted by the edge beams was 90% of
the applied load due to flexure of the floor slab. The loading by the jacks was converted to an
equivalent uniform loading by dividing the jack load by an area of 9.95 m2 for the internal beams
(loaded area of 5.24 m x 1.92 m).

65
For loading by two point loads through the secondary beams, the equivalent uniform loading that
the primary beam could resist was calculated as 31.5 kN/m2, having subtracted the self-weight of
the slab and the loading rig, which is estimated as 3.3 kN/m2. This uniform loading corresponds to
a jack load of 314 kN per beam. The A142 mesh reinforcement was calculated to be just sufficient
to transfer the longitudinal shear force. Therefore, longitudinal failure though splitting of the
concrete was at its design limit although there was no evidence that the shear resistance of the
shear connectors was affected.

Loading was applied in 4 initial cycles and the load-displacement curves for the first 3 cycles of
loading up to an equivalent uniform loading of 20 kN/m2 shows that the response is essentially
elastic. The load-displacement curve for the secondary beams is shown in Figure 5.7 and failure
corresponded to an absolute deflection of approximately span/70. The load was then increased
incrementally to the eventual failure load of 46 kN/m 2, which corresponded to a jack load of 458
kN per beam. The total equivalent uniform failure load was therefore 49 kN/m2 including the self-
weight of the slab.

Figure 5.7 Deflection of secondary beams close to failure

Failure occurred in the secondary beams despite them being theoretically 14% stronger than the
primary beams. The failure mode was apparently due to longitudinal shear failure caused by the
low degree of shear connection. The connections of the secondary beams to the primary beams did
not transmit high moments as seen from the deflected shape in Figure 7.7.

The load-displacement curve for the primary beam is shown in Figure 7.9. The deflection of the
primary beam at the failure load was 22mm, or span/163, and it is apparent that the primary beam
was not close to failure. The load-displacement curve for edge primary beam is shown in Figure
7.7, which is still elastic. The average of the deflection of the internal and edge primary beams was
14 mm at failure, and so the net deflection of the secondary beam at failure was 68-14 = 54 mm
(=span/96).

The measured deflections of the adjacent edge beams were approximately 10% of that of the
internal beam. It is estimated that the primary beam was subject to 80% of the load applied to the
secondary beams, which corresponded to a point load of 0.8 x 458 = 366 kN. This equates to a
moment of 366 x 3.6/4 = 329 kNm (equivalent uniform load of 30 kN/m2). The moment acting on
the secondary beams was 0.5 x 329x1.92 = 316 kNm plus a moment of 25 kNm due to the self-
weight of the slab, beams and loading frame, which equals 341 kNm.

66
50
45
40
35
Floor Load (kN/m2)

30
25
20
15
10 LVDT-2
5
0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75
Beam mid-span deflection (mm)

Figure 5.8 Load-displacement graph for internal secondary beams up to failure

Figure 5.9 Load-displacement graph for internal primary beam up to failure

The theoretical bending resistances and degrees of shear connection for the secondary beams and
primary beam are presented in Table 5.1. The unity factor for the secondary beams was 341/358 =
0.95, which is consistent with normal level of acceptability for plastic design for a low degree of
shear connection of 26%. The unity factor for the primary beams was 1.14, which indicated that
the beam had a 14% reserve of safety despite being designed to fail first.

Table 5.1 Summary for test on primary beam with loaded internal secondary beams allowing for
10% load spread to each of the edge beams

Beam Measured material properties


 Equivalent MRd Test failure UF in Reinforcement
uniform load (kNm) moment inc. bending As,provided/As,req
(kN/m2) self wt.

Primary Internal 0.35 315 329 + 30 1.14 0.99


0.8 x 46 + 3 self weight
= 40 kN/m2
Secondary 0.26 358 337 + 22 0.95 >>1
Internal self weight
 is the degree of shear connection of the edge beams based on a shear connector resistance of
65kN for the primary beams and 61 kN for the secondary beams.

67
In this table, UF is the unity factor in bending (Mapplied/MRd) allowing for 20% load transfer from the
internal secondary beam to the edge beams.

The moment transferred to the beam to column connections at the primary beams was small as
evidenced by the negligible forces in the tie members between the columns.

5.3.2 Tests on edge beams


The load test on the edge beams each subject to two point loads is shown in Figure 5.10. The first
test was on the edge beam without U bars. To calculate the loaded area acting on the edge beam,
the loaded width of the slab was calculated based on half the spacing between the centre-line of
the beams (=1.89 m/2) plus a slab edge distance of 0.1 m. The loading by the jacks was converted
to an equivalent uniform loading by dividing the jack load by 5.5 m2 (loaded area of 5.24 m x
1.05 m).

Figure 5.10 View of load test on edge beam

The load deflection curve is shown in Table 5.1. It is apparent that the behaviour was essentially
elastic up to an equivalent uniform loading of 30 kN/m2 and the residual deflection on unloading
was only 2mm. Cracking due to longitudinal shear failure was evident at an equivalent uniform
loading of 37 kN/m2, as shown in Table 5.2.

The failure load occurred at an equivalent uniform loading of 60 kN/m 2 that corresponded to a
point load of 165 kN. The applied load acting on the edge beams was then reduced by 10% to take
account of the transverse stiffness of the slab, which transfers load to the internal beam.

The shear connector resistance for edge beams without U bars was taken conservatively as:

PRd,edge = 0.5 PRd (1 +e/(2hsc)) = 61 x0.5 x (1+100/(2 x 95 )) = 61 x 0.76 = 46 kN

where e is the distance of the centre of the shear connector from the slab edge =100mm

hsc is the as-welded height of a shear connector = 100-5 = 95mm

PRd is the design resistance of a shear connector for an internal beam = 61 kN

The test bending resistances and theoretical bending resistance are presented in Table 5.2.

The model factor (Mtest/ Mcalc) was 1.02 based on the estimated shear connector resistance of 46
kN, which is 76% of the case with U bars .when using the above formula. The test bending
resistance was consistent with development of a shear connector resistance of 50 kN without U
bars.

68
Table 5.2 Summary for test on edge beams with and without U bars

Secondary  Equivalent MRd Test Model Reinforcement


edge beam uniform load (kNm) moment, factor in As,provided/As,req
(kN/m2) Mtest (kNm) bending
Beam with no 0.32 0.9 x 60 + 3 292 285x+12 1.02 x <1
U bars+ (self weight)
Beam with 0.52 0.9 x 63 + 3 313 299 +12 1.0 >1
U bars
x Shear connector resistance without U bars is taken as 76% of that with U bars

The load resisted by the edge beams is reduced by 10% due to the transverse stiffness of the slab.

65
60
55
50
45
40
Floor Load (kN/m2)

35
30
25
20 LVDT-8
15
10
5
0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75
Side Secondary Beam Mid-span Deflection (mm)

Figure 5.11 Load -displacement graph for the edge beam test without U bars

Figure 5.12 Edge beam test without U bars at an Figure 5.13 Deflected shape of edge beam
equivalent loading of 37 kN/m2 test with U bars close to failure

The second test was on the edge beams with U bars was subject to the same loading system so
that the failure loads could be directly compared. The edge beam at failure is shown in Figure 5.14.
The load-displacement graph for this case is shown in Figure 5.15.

69
70
65
60
55
50
45
40
Floor Load (kN/m2)

35
30
25
20
15 LVDT-6
10
5
0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
Side Secondary Beam Mid-span Deflection (mm)

Figure 5.14 Load -displacement graph for the edge beam test with U bars

It is apparent that the beam was slightly stiffer than for the same case without U bars (compare a
deflection of 23mm and 20mm at an equivalent loading of 40 kN/m2). This was probably due to the
evidence of a lower level of cracking in this test which reduced the longitudinal slip. Also, the
residual deflection on unloading to 40 kN/m2 was 3.8 mm in comparison to 4.8 mm for the case
without U bars.

The failure load for the edge beam with U bars corresponded to an equivalent uniform loading of 63
kN/m2 plus the self-weight of the beam and slab (3.3 kN/m2). When reduced by 10% to allow for
the load transfer to the internal beam, the failure load was 60 kN/m2. The bending moment at
failure was equal to the plastic bending resistance for a shear connector resistance of 61 kN, which
is the same for an internal secondary beam.

5.4 Results of robustness test


The robustness test involved removal of support to an edge column, as shown in Figure 5.15. The
measured deflection of the central edge column on removal its support was 36 mm when subject to
load of 3 kN/m2, due to the self-weight of the slab and beams. The equivalent load acting on the
missing column was 33 kN due to the floor self-weight. The deflection corresponded to an
inclination of 1 in 145 for the edge beam, which is small.

Figure 5.15 Deflection of edge beam due to missing support to central column

The point load was applied at mid-span of the edge beams. The load acting on each beam was
divided by an area of 5.5m2 in order to obtain the equivalent uniform loading, as shown in Figure
5.16. The maximum displacement of 190mm was reached for an equivalent uniform loading of
13.8 kN/m2 in addition to the self-weight, which corresponded to a point load of 75 kN per jack
acting on each beam. This corresponded to a notional load of 75 kN acting on the central edge
column with its missing support. The maximum displacement corresponds to an inclination of 1 in
27 or about 2o for the edge beam.
70
It is apparent that the maximum load capacity of the floor system had not been reached at the
maximum displacement of 190mm, and a further increase of 20 to 30% in loading might have
been possible. This would correspond to a missing column load of 33 +100 = 133 kN.

Figure 5.16 Deflection of column with a missing support subject to an equivalent uniform loading
acting on the edge beams

Significant cracking in the slab and plastic deformation of the connections was observed at the end
of the test. Under the self-weight of the slab, a strain of about 140 μ strain was recorded in the
bottom flange which corresponds to a tensile stress of 29 N/mm 2. For the subsequent loading, a
strain of about 170 μ strain was recorded in the top flange and 200 μ strain in the bottom flange.
This indicates a tensile stress of 39 N/mm2 in the cross-section and a tension force of 179 kN.

The total applied load on the missing column is 33 + 75 = 108 kN. For an edge beam inclination of
1 in 27, the catenary force in the edge beams could reach a maximum of 108x27/2 = 1458 kN.
This shows that the catenary action in the test was about 12% of this maximum value, and the
remaining resistance was provided by the bending resistance of the connections and membrane
forces in the slab.

5.5 Modelling of floor plate test

5.5.1 FE model description


The ABAQUS software package was used to develop a non-linear FEM to understand the structural
behaviour of the composite floor plate. Due to the large size of the floor plate, some members
were simplified to minimise the ABAQUS running time. The following element types were adopted:
Three-dimensional 8-node solid elements (C3D8) were adopted for the concrete slab. The shell
elements (S4R) were adopted for the decking profile. Beam elements were adopted for steel
beams, columns, shear connectors and load-spreading beams. Beam (truss) elements were used
for reinforcement mesh.

The joints were simplified to a “tie”. Therefore, an “MPC tie” was adopted for the beam to beam
and beam to column connections. As the decking was welded to the beam via the shear connector
root, a “tie” was used to fix the decking to steel beam by selected points. “Contact” was adopted to
simulate the interaction between decking profile and the concrete. In the normal direction, a “hard”
condition was adopted and in the tangential direction, a “penalty” condition with a friction factor of
0.3 was adopted. The contact interaction between the slab and steel beam was not taken into
consideration. Instead the shear connecter was embedded in the concrete slab.

The steel behaviour for the beams, columns and shear connectors was modelled using a tri-linear
stress-strain relationship. The concrete stress-strain behaviour was modelled using the concrete
damaged plasticity model available in ABAQUS. The steel behaviour for the reinforcement and
decking profile was modelled using the perfect elasto-plastic stress-strain relationship. The shear
connectors were assumed to have a nominal yield stress of 450 MPa and an ultimate strength of
520 MPa .

71
5.5.2 Modelling results for jack loads applied to the internal secondary beam
Figure 5.17 shows the model with jack loads applied to the internal secondary beams. Figure 5.18
compares the displacements at the mid-span of beams in the FE prediction and test results. The
deflections in the FE model are very close to those measured. For the internal primary beam, the
initial part of the loading was modelled accurately but the elasto-plastic part is not, which suggests
that membrane action in the floor plate model was more significant than in the test.

For the internal secondary beam, both the initial part of the loading and the elasto-plastic part
were modelled accurately. Based on the measured and modelled deflections of the secondary edge
beams, and allowing for their difference in size to the internal beam, it was determined that 10%
of the applied load was transferred to each edge beam due to the flexural stiffness of the slab.

Figure 5.17 Deformation of floor plate with loads on the internal secondary beams

Figure 5.18 Comparison of FE modelling against test results for internal beam test

72
5.5.3 Modelling results for loads applied to the edge beam
Figure 5.19 shows the deformation of the composite floor plate and stress distribution in the steel
beams. Figure 5.20 compares the displacements at the mid-span of beams in the FE prediction
with the test results. The vertical deflections are very close in the elastic part but the predicted
failure load was about 30% higher. This is attributed to the difficulty in modelling the behaviour of
the shear connectors when affected by the slab edge, as they are more deformable than in push
tests.

At failure, the displacement of the internal secondary beam was about 5% of that of the edge
beam. Allowing for the difference in size of these beams, it is concluded that about 8% of the
applied load is transferred from the loaded edge beam to the internal beam due to the flexural
stiffness of the slab. An estimate of 10% was used in the comparison with the plastic bending
resistance in Section 7.3.1.

Figure 5.19 Deformation of composite plate with loads on the edge beams

Figure 5.20 Comparison of FE model prediction the tests on edge beams

5.5.4 Concluding remarks for WP4


The following conclusions are made from the floor plate tests:
 The internal primary beam was found to be stronger than the plastic bending resistance despite
having slightly less transverse reinforcement than required by Eurocode 4. The failure load was
14% higher than that calculated using a shear connector resistance of 65 kN, which was
calculated by reducing the push test result of 89 kN according to the measured concrete
strengths. The internal secondary beams failed at an equivalent uniform loading of 49 kN/m2,
which when reduced by 20% as noted above, gave a moment of 95% of the plastic bending
resistance.

73
 For the load tests on the edge secondary beams, the edge beam with U bars around the shear
connectors failed at an equivalent uniform loading of 66 kN/m2 (including the self-weight), and
the beam without U bars failed at an equivalent uniform loading of 63 kN/m2 (5% less).
Considering a 10% transfer of load from the edge beam to the internal beam, the edge beam
with U bars failed at a moment exactly equal to the plastic bending resistance.

 For the test without U bars, the shear connector resistance corresponding to the plastic bending
resistance was calculated as being 17% less than for the case with U bars. This corresponds to
a longitudinal shear resistance of 50 kN. Based on this result, the shear connector resistance for
edge beams without U bars may be taken conservatively as:

PRd,edge = 0.5 PRd (1 +e/(2hsc)) for e ≤ 2 hsc

where:

e is the distance of the centre of the shear connector from the slab edge

hsc is the as-welded height of a shear connector

PRd is the design resistance of a shear connector for an internal beam

 In terms of deflections, it was apparent from the tests on the internal secondary beams that the
effects of slip on deflections were less than calculated for a load of 10 kN/m2. This is probably
due to the continuity effects of the mesh reinforcement in the slab. The measured deflection of
the primary beam was higher than calculated for a load of 20 kN/m2,which shows that the
stiffness of the connections to the columns is relatively small.

 For the edge beams subject to an equivalent uniform load of 20 kN/m2, the measured
deflections were 7.4 mm with U bars and 8 mm without U bars compared to a theoretical
deflection of 6.2 mm including the effects of slip. This shows that for edge beams, the effects of
slip are greater for higher load levels. Comparisons were very close for lower load levels. The
effect of the lack of U bars was relatively small.

 For the robustness test, it is concluded that the floor structure can resist an equivalent load of
108 kN in the missing column, which when divided by the supported floor area of 10.5 m2,
corresponds to a floor loading of about 10 kN/m2, including the slab self-weight. This
demonstrates the robustness of composite floors to damage of the columns.

 The Abaqus FE modelling results are close to the test results. Thus, the beam elements may be
suitable for modelling the steel section. Furthermore, the concrete damaged plasticity model
provided sufficiently good results to define the response of the concrete slab.

5.5.5 Exploitation and impact of the research results


The results are directly applicable in design as it was shown that a modification factor may be
made to the shear resistance of the shear connectors at edge beams without U bars. It was also
shown that the discontinuity in the steel decking at the primary beams did not lead to premature
failure by splitting which is a practical case in design.

The robustness test showed that the load from the missing column equal to the self-weight of the
slab plus the accidental load acting on the floor may be supported by the floor plate assisted by
tying action at the columns.

74
6 WP5: Numerical models for composite beam analysis using shear
connector properties

6.1 Objectives of WP5


The objective of this Work Package was to develop numerical models that reflect the realistic
behaviour of the tests and to use the numerical models to analyse composite beams of various
proportions, such as:

 Symmetric and asymmetric beams with asymmetry of flange areas up to 3 to 1.


 Cellular beams with regular circular openings.
 Propped and unpropped beams in the construction stage.
 Ductile and semi-ductile shear connectors, as defined in WP1 ( a new category of super-ductile
shear connectors with a limiting slip of 10 mm was identified in the course of this research and
the performance of composite beams with these shear connectors which was added to the
objectives).
 Different loading patterns, such as point loaded beams.
 The models developed in this WP were correlated with the test results taking account of:
 Strain hardening in the steel.
 Concrete crushing and un-loading at high strains.
 Unloading shear force-deformation relationship for the shear connectors.
 Frictional effects due to direct loading through the slab.
 Material and geometrical non-linearity including the cracking and crushing of the
concrete.
 The steel decking and reinforcing mesh.
The models used were based on ANSYS and ABAQUS finite element (FE) software packages and
the data-files were exchangeable among the partners. The models used for the analysis of beams
used appropriate load-slip relationships for the shear connectors. More detailed models were
required for the local failure at the shear connectors but these are not appropriate for modelling
whole beams because of the excessive computing time. This concrete cracking model extended to
the effects of concrete splitting local to the shear connectors.

6.1.1 Comparison of initially planned activities and work accomplished


The analysis of the composite beams required use of various finite element packages and it was
agreed that the work would be distributed according to the available packages.

6.2 Description and impact of the research results

6.2.1 Shear connector failure model


For shear connectors in solid slabs, according to Lungershausen, a distinction is made between the
four contributions to the load-bearing behaviour. This depends on the slip “w” (see Figure 6.1).
Initial slip develops flat struts (𝛽 ≤ 35°) that are supported by the welding collar at the foot of the
headed stud (A). Increasing the load leads to concrete damage at the foot of the headed stud.
Associated with this is a redistribution of forces to the shaft, which is stressed by bending forces
from this point (B). This leads to a plastic deformation of the shaft of the stud.

Figure 6.1 Load bearing behaviour of a headed stud in a solid slab, Lungershausen

The restrained deflection leads to a tension force in the stud, which is in equilibrium with the
concrete compression force below the head = (C). Increasing the load further leads to friction
forces between concrete cone and surface of the steel beam (D). The higher tension force C is, the
75
lower is the bending stress B of the shaft. Failure of the headed stud may occur either because of
concrete failure or shaft shearing of the stud.

For shear connectors used with composite slabs, the total shear force is resisted by bending of the
shaft of the stud (Part B). Increasing the load leads to concrete failure in front of the shaft (Figure
6.2). A sufficient embedment depth of the headed stud ensures that two plastic hinges are
developed in the stud when reaching the maximum bearing capacity.

Increasing slip leads to a vertical force in the concrete below the head of the stud. This leads to
tension stresses in the shaft of the stud and compression in the concrete, which react against the
flange of the steel beam. The horizontal component of the tension force in the shaft of the stud
affects the load in C. If the embedment depth of the headed stud is not sufficient, the second
(upper) plastic hinge may not be developed. Then, the load-deflection curve reduces after reaching
the first maximum load P1u.

Figure 6.2 Behaviour of headed studs in profiled slabs according to Lungershausen

The observed failure mechanisms in the push tests were:

 Stud failure – Shearing of the stud shank


 Concrete pull-out – 2 plastic hinges in the stud shank generally observed after the second
maximum load
 Rib punch-through – Concrete break-out because of concrete failure in front of the shank and
punch through the decking.

6.2.2 Finite element modelling of shear connectors


The work on modelling of shear connectors was carried out using the finite element software,
ABAQUS Explicit. All parametric studies were carried out using the ABAQUS Version 6.13-5. The
explicit solver is suitable for dynamic or high non-linear problems. An explicit calculation does not
need to solve the whole system of equations which leads to faster calculations. The first step was
to model the test V 1-03 on a solid slab which was the basis for all further test series (not reported
but given in the Deliverable to WP5.1).

The double symmetry of the specimen made it possible to model only a quarter of the push-out
specimen shown in Figure 6.3. This saved computing time because the complicated geometry of
the decking lead to a very fine meshing in the trough. The concrete slab, steel beam, tension bars
and headed studs were meshed using hexahedral elements C3D8R that is a universal applicable
linear element with 8 nodes. Reduced integration leads to shorter calculation times but may also
leads to a reduced stiffness of the element, called “hour-glassing”.

Due to the geometry of the trough of the CP 60 decking, a recess was formed in the concrete. To
mesh this wedge properly, a C3D6R element was used. This is a linear element with 6 nodes and a
shape of a triangular prism. The end plate of the steel beam (load introduction point) and the
contact surface of the concrete were modelled using R3D4 elements, a three-dimensional rigid
element with 4 nodes. It is infinitely stiff and cannot be deformed.

The reinforcement bars were modelled using B31 elements. These are three-dimensional beam
elements. The profiled decking consists of S4R elements. This is a shell element with four nodes
and reduced integration.

76
Concrete
Steel beam

Headed stud

Tension bar

Figure 6.3 ABAQUS model for shear connectors in composite slabs

All contacts are defined in tangential and normal direction. The behaviour in the normal direction
was defined as “hard”-contact overclosure relationship. This led to a very low penetration of the
two surfaces. In the tangential direction, friction between the surfaces was assumed. Actually the
connection between profiled decking and concrete should be defined by a contact but this led to
unrealistic dynamic response of the decking during the explicit calculation. Therefore, the
connection was performed as a tie-connection by which the deflection of the decking across to the
direction of the force was prevented and could not be compared to the real test results.

The behaviour of steel was given by a bi-linear stress-strain relationship. Modelling of the concrete
was used the concrete damaged plasticity model of ABAQUS. This model takes into consideration
concrete compression and tension failure. It was not able to produce real concrete cracks but
rather considered cracks by a stiffness reduction as soon as crack strains were exceeded. This is
called “smeared cracking”. This model requires some additional material parameters, as shown in
Table 6.1.

Table 6.1 Material parameters concrete damaged plasticity

Dilation angle Eccentricity fb0/fc0 K Viscosity


40 0,1 1.16 0.6666 0

Modelling the concrete under compression was carried out to EN 1992-1-1: Eurocode 2. Firstly,
the concrete compression strength was determined by three identical 150 mm cubes. The cylinder
value was obtained by multiplying the cube strength by 0.8. The mean value was calculated as the
characteristic value +8 N/mm2. This compression strength was used for the stress-strain curve, as
in Table 6.1: The elastic part of the strains εel was subtracted from the total strains.

Comparative calculations showed that Eurocode 2 overestimates the concrete strain for the
maximum concrete compression strength. Therefore, the strain corresponding to the maximum
concrete strength was reduced to 0.88 ‰, which leads to earlier concrete failure. The
characteristic load drop off after having reached a first maximum could not be simulated. This
means that the -ε-behaviour was shifted left and compressed at the same time, which led to an
earlier failure of the concrete, as shown in Figure 6.4. The short term elastic modulus was taken as
E = 33.8 kN/mm².

77
Stress-Strain Behaviour
Spannungs-Dehnungsverhalten Concrete Damage
Betondamage
45.00 0.7

40.00
0.6

Compression Damage Parameter


35.00
0.5
Spannung [N/mm²]

30.00

Stre

mm
[N/

crete
Dam

mete
Para
Con
ss

²]

age

r
25.00 0.4

20.00 0.3

15.00
0.2
10.00
0.1
5.00

0.00 0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
plastischePlastic strain
Dehnung [‰] [‰] Plastic strain
plastische [‰][‰]
Dehnung

Figure 6.4 Concrete behaviour in compression

In addition to the concrete behaviour under compression, the concrete damage behaviour had to
be determined. The compression damage parameter is zero until the maximum concrete
compression strength is reached. The graph shows the concrete damage depending on the plastic
strain. The tensile strength is determined according to CEB-FIP Model Code.

The post-cracking behaviour is characterized by a bi-linear relationship. The tensile strength fctm
corresponds to the compression strength to Eurocode 2. For fctm = 3.2 N/mm², values are
calculated in Figure 6.5(a). ABAQUS requires the use of a tension damage relationship of the form
shown in Figure 6.5(b).

Figure 6.5 Stress-crack width and tension damage relationship for concrete

6.2.3 Verification of the FE model


The determined material properties, contact options and the FE meshing lead to the red load
deflection curve, see Figure 6.6. The initial stiffness fits well. A first local maximum was reached
before the load decreased slightly and then increased again. The first local maximum was reached
at 4.7 mm in the FE simulation, which is between the real measured values 2.5 mm and 5.1 mm
for the local maximum. The load reached at the local maximum is 87 kN /stud in the FE analysis,
which corresponded well with the results of test 1-03-2.

78
V1-03
140

Force/stud [kN] 120

100
Kraft pro Kopfbolzen [kN]

80

Versuch1-03-1-s
Versuch1-03-2-s

60 Versuch1-03-3-s
Abaqus

40

20

Slip [mm]

0
0 2 4 6 8 10 12 14 16 18 20
Weg [mm]

Figure 6.6 Comparison of test results and ABAQUS results for test1-03-2

Figure 6.7 shows the good agreement between calculated and real deflection of the headed stud.
In both cases, two plastic hinges were developed, at first at the foot of the stud, and then in the
middle of the shank. Above the second plastic hinge, the shank remains straight. The height of the
second plastic hinge corresponds with the height of the decking. The welding collar remained
intact, there was no shear failure of the headed stud. In addition, the FE calculation showed that
the concrete and stud separated from each other on the non-loaded side, while the loaded side was
in compression and led to buckling of the decking and finally to rib punch through failure.

Figure 6.7 Deformation of the headed studs within the concrete slab

To make concrete ’cracks’ visible, the FEA shows those regions that have reached the plastic
concrete strains. The grey regions have already exceeded the maximum strain. A comparison of
test results and FEA shows a horizontal crack in both cases between the positions of the headed
studs. In addition, transverse cracks between the lower headed stud and the upper edge of the
recess can be observed, see Figure 6.8.

The horizontal crack occurred before the transverse cracks and resulted from the different load
distribution among the studs at the beginning of the push test. The lower stud was activated before
the upper stud. This resulted in different concrete stresses acting on the shear connectors, which
led to this kind of cracking. The transverse cracks occurred due to the load transfer to the support.

79
Above the recess, tension forces acted on the concrete and the reinforcement. The internal corners
of the recess cause concentrated stresses, which also lead to cracks.

(a) Plastic strains in the concrete (b) Strains and cracks on the concrete surface
Figure 6.8 Plastic strains of the concrete slab obtained from the FE model

To analyse the failure mechanisms, the composite slabs were opened after the tests. In addition to
the already described rib punch through failure, a second failure mode called concrete pull-out was
identified. This failure mode is characterised by a concrete cone which was broken out below the
head of the stud, see Figure 6.9. The concrete damaged plasticity model of ABAQUS made it
possible to show those areas of the concrete that were severely damaged. Comparing the breakout
cones shows a good agreement of FEA and the real tests. The cone of the lower stud (the right one
in the figure) was clearly stronger damaged than the other one. This confirmed the assumption
that the lower stud resisted more load at the beginning of the push test.

Figure 6.9 Concrete damage and concrete pull-out failure of test 1-03-3 compared to ABAQUS
model

6.2.4 Parametric study


ABAQUS was used to investigate the influence of several parameters on the load-slip behavior, as
follows:

 Diameter of the headed stud


 Height of the headed stud
 Concrete strength
 Influence of reinforcement layers
 Recess of the contact surface
Other cases are presented in the Deliverable to WP5.1. To be able to compare the results, only one
parameter was changed at a time. Table 6.2 shows the effect of changing the shear connector
diameter for the CF60 deck profile in comparison to the expected change according to Eurocode 4.

80
Table 6.3 shows the effect of changing the shear connector height for a 22mm diameter shear
connector and the CF60 deck profile which leads to an increase in embedment depth. This is given
by the ratio hsc/hp (CP60 decking has hp=58 mm).

Table 6.2 Change of the stud shear resistance with various parameters

Diameter / height of (dsc /22)² Expected Change in Calculated change


headed studs [mm] change (EC4) real test (ABAQUS)

16x125 0.53 -47% -36%

19x125 0.74 -25% -40% -30%

22x125 1.0

25x125 1.29 +29% +20%

Table 6.3 Change of the stud shear resistance with various parameters (initial value D22x125)

Diameter / hsc/hp kt (EC4) k Expected Change in Calculated


height of (Konrad) change real test change
headed studs (EC4) compared (ABAQUS)
[mm] to 1-03

22x75 1.29 0,284 0,375 -62% -23% -16%

22x90 1.55 0,54 0,45 -28% -7% -8%

22x100 1.75 0,71 0,474 -5% -2% -3,5%

22x125 2.11 0,75 limit 0,485

Eurocode 4 assumes that the shear resistance is proportional to √fck. in case of concrete failure.
According to Konrad, the shear resistance is proportional to fck2/3. The results of the different
calculations can be found in Table 6.4. One pair of real tests used the same configuration but
different concrete grades C30/37 and C40/50. The load increase of the real tests was + 11.5%
from a C30/37 to a C40/50 grade.

Table 6.4 Change of the stud shear resistance (initial value C30/37)

Concrete grade ~√𝒇𝒄𝒌 ~𝒇𝒄𝒌 𝟐 𝟑 Expected Expected Calculated
(EC4) (Konrad) change change change
(EC4) (Konrad) (ABAQUS)

C25/30 5,00 8,55 -9% -11% -8%

C30/37 5,48 9,65

C35/45 5,90 10,70 +8% +11% +18,0%

C40/50 6,32 11,70 +15% +21% +40%

Comparing the predictions of the equations and the real test results to the FEA, it can be seen that
the ABAQUS analysis over-estimates the shear resistance for concrete grades higher than C35/45.
This may be caused because of lack of information on higher concrete grades.

Other analyses were carried out for single and double layers of reinforcement. The tests showed
that a second layer of reinforcement positioned directly on top of the decking increases the ductility
of the system, but has little influence on the shear resistance. The FEA confirms the test results for
this case. The shear resistance was is 83 kN when using only the top layer of reinforcement. The
second (bottom) layer of reinforcement leads to a load increase of about 6% and an increase of
ductility of more than 64%.

A recess at the support is often used in push tests. To evaluate the effect of the recess, the push
test 1-03 was modelled with and without the recess. It was found that that an additional recess
had no influence on the total load-deflection behaviour of the push-out test. An additional crack
was noticed above the edge of the recess, which implies that the load spreading to the base of the
concrete slab is caused by the recess.
81
6.2.5 Final remarks on WP5.1
This deliverable shows that Finite Element Analyses may be used to model accurately the load-
deflection behaviour of push-out tests with composite slabs. Several parametric studies
investigated the influence of selected parameters on the shear resistance of the shear connectors,
as follows:
 The developed model shows a good correlation with the performed tests when using the
concrete damaged plasticity model for the material properties of the concrete slab.
 The shear resistance of headed studs increases when using larger stud diameters. However, the
load increase is not proportional to dsc2 as given in EC4.
 The shear resistance of headed studs increased also when using higher studs. The embedment
depth hsc/hp was found to be an important parameter in the failure mode.
 Regarding the influence of the concrete strength, the formulae given in EC4 fit well to the test
results. For higher grade concretes, more push tests are needed to calibrate the FE model.
 The influence of the reinforcement is not covered by EC4. Adding a second layer of
reinforcement placed directly above the decking leads to an increase of ductility, but only to a
negligible change of the shear resistance.
 The recess at the support of the specimen has nearly no influence on the shear resistance, but
the FEA shows a clearer load distribution when performing the test with recess.
6.3 WP 5.2 Modelling of composite beams
The two long span beam test in WP3 were modelled in the scope of this WP in order to verify the
accuracy of the finite element models using both ABAQUS and ANSYS. An idealised relationship
between force and slip was adopted for the shear connectors in the FE models based on the results
in WP1 using CF80 decking. This is based on stiffness of 70 kN/mm for single shear connectors per
deck rib, a shear resistance of 72 kN and a limiting slip of 5mm. A declining part of the curve was
added to a slip of 10mm. The concrete and steel properties are shown in Figure 6.10.

In the ABAQUS modelling, two FE models were investigated. FE Model 1 used ABAQUS Explicit
method and 3D solid elements for the steel beam, concrete and shear connectors and also included
the steel decking, and it proved to be unstable in its output and complex in computing time. FE
Model 2 removed the steel decking and used shell elements for the flanges and web of the steel
beam that were connected to the slab by springs representing the shear connectors. This proved to
be more accurate and simpler.

(a) Steel properties (b) Concrete properties to EC2

Figure 6.10 Steel and concrete properties used in the FE models

The load versus mid-span deflection and load versus end-slip relationships for FE Model 2 are
shown in Figure 6.11, along with FE Model 1 and the test results. FE Model 2 was more flexible
than the test results and also FE Model 1. Furthermore, the maximum mid-span deflection and
maximum end-slips in FE Model 2 were almost twice as high as those in the tests. However, the
load progression in FE Model 2 follows the test results more closely above 12 kN/m2 and predicts
the failure load.

82
(a) Load versus deflection according to FE (b) Load versus end slip according to FE
prediction prediction

Figure 6.11 Comparison of FE model and 15m span cellular beam test in WP3

To analyse the 11.2 m span asymmetric beam, the ABAQUS static general method was adopted.
3D solid element was adopted for shear studs, steel beam and concrete slab, No reinforcement and
no steel decking were included. Two concrete grades were adopted in the parametric study
according to the measured concrete properties of the tested specimen.

Figure 6.12 shows the comparisons of the tests and numerical modelling for this beam. It can be
seen that the prediction of beam mid-span deflection agreed well with the test results before the
load reached 15 kN/m2. This is possibly due to the fact that the ABAQUS model did not capture the
concrete damage well. For the end-slip and the strains in the steel beam, the numerical modelling
gave a good agreement with the test results up to a load of 15 kN/m2.

(a) Load versus deflection according to FE (b) Load versus end slip according to FE
prediction prediction

Figure 6.12 Prediction of the performance of the 11m span asymmetric beam test using ABAQUS

Propped versus unpropped beams


It is important to clarify the definition of un-propped construction and what are the differences with
propped construction in terms of the shear connection requirements. Where the beam is
constructed as un-propped, the self-weight is applied to the steel beam alone, and so no load is
applied to the shear connectors. This leads to a potential reduction in the minimum degree of shear
connection of un-propped beams. However, the benefit of un-propped construction in terms of the
minimum degree of shear connection reduces as the imposed load increases as a proportion of the
total load on the beam. This is an important concept, as follows:
 The plastic bending resistance of the beam is independent of the order of loading and whether
loads are applied to the steel beam in construction.

 However, the minimum degree of shear connection will vary depending on the proportion of the
total loading that is imposed and acts on the composite beam.

83
 Previous guidance defined minimum degree of shear connection limits for un-propped
construction that only apply to beams subject to a factored imposed uniform loading of not
greater than 9 kN/m2 (NCCI). That definition has been refined and complemented by an upper
limit on the moment due to factored imposed load as a percentage of the moment due to total
factored loading. This revised definition means that other types of loading can also be
considered.

 Studies in WP 6.4 showed that uniform loading lead to the greatest slip when compared to
beams with point loads at the third points of the span and for beam with central point loads.
Therefore, the minimum degree of shear connection rules derived for uniform loading can be
used for point load cases.

6.4 WP5.3 Modelling of cellular beams with low degrees of shear connection
Cellular beams with multiple circular openings were analysed using both ABAQUS and ANSYS FE
codes, and were validated against the 15.3m test in WP3. Various parametric analyses were
performed with the objective of comparing the minimum degree of shear connection requirements
for cellular beams with the equivalent for solid web beams. These results were also compared to a
formula for ηmin for cellular beams, which is a modified version of the EN 1994-1-1 formulae.

The ABAQUS 6.13 package offers a range of tools to represent the behaviour of steel-concrete
composite beams subjected to static loading. Steel beams and reinforced concrete slabs may be
modelled, for example, by use of 3-D brick elements. Modelling of the shear connection is
described in Figure 6.13:

Rigid plane tied to slabs (2X)

MPC2 (Beam)

Nonlinear spring
MPC1 (Beam)
Slab-profile contact
Top flange
(no friction)

(a) View of FE model (b) Spring model

Figure 6.13 Description of the longitudinal shear connection model in ABAQUS

6.4.1 Cellular beam model in ABAQUS


The cellular beams were modelled with the following properties:

 Shear connectors modelled using bi-linear springs (Fmax = 70 kN ; elastic slip, delast = 1 mm) ;
 Non-linear concrete material law from EN 1992-1-1 using Concrete Damaged Plasticity Model
(with fcc = 25 MPa ; fct = 3 MPa ; Dilation Angle = 40°, Eccentricity = 0.1, fb0/fc0 = 1.16, K =
0.666, Viscosity Parameter = 0, Tension stiffening = 1.2 mm ; Linear evolution of tension and
compression damages beyond peak stresses);
 Bilinear stress-strain diagrams for steel profile (fy = 410 MPa ; E = 210,000 MPa) and
reinforcing bars (fy = 500 MPa ; E = 210,000 MPa) ;
The failure mode obtained was a bending failure after development of high stresses in the bottom
flange at mid-span (see Figure 6.14). The correlation between the test deflection and F.E. results is
good (see Figure 6.14 (a)). Mid-span deflections were slightly over-estimated for loading higher
than 8 kN/m2.

At the end of the test, the FE model resisted a loading slightly higher than 18 kN/m2 but it should
be noted that the difference of behaviour occurred during an unloading-reloading cycle in the test.
The correlation between end slips measured experimentally (average value between LVDT_10 and
LVDT_11) and predicted by the FE model was very good (Figure 6.14 (b)).

84
(a) Stress distribution in cellular beam around the openings for a (b) Stress range
case of low degree of shear connection

Figure 6.14 Distribution of Von Mises stresses in the beam at failure

WP3 test - Load-deflection diagram WP3 test - Load-slip diagram


20 20

18 18

16 16
Distributed variable load (kN/m2)

14 14

Total variable load (kN/m2)


12 12

10 10
Experimental Experimental
8 8
Model 27 Model 27
6 6

4 4

2 2

0 0
0 50 100 150 200 250 0 2 4 6 8 10 12 14 16 18
Mid-span vertical deflection (mm) Beam end-slip (mm)

(a) Comparison of load versus deflection (b) Comparison of load versus slip

Figure 6.15 Comparison between mid-span deflections for the 15.3m span cellular beam test and
obtained using ABAQUS

6.4.2 Cellular beam model in ANSYS


A second F.E. model of the cellular beam was created using ANSYS in order to compare the results
with the ABAQUS model. This FE model was built using shell elements for the steel section and
solid elements for the slab. The shear connectors were modelled using non-linear spring elements.
The slab ribs were also modelled but they were given reduced material properties (the Young’s
modulus was set to Ec/10) in comparison to the concrete flange. This was done to avoid over-
estimating the system stiffness, since the contribution from the deck ribs is already accounted for
in the load-slip characteristics which are input to the spring stiffness.

The FE model was validated against the 15.3 m span cellular composite beam test of WP3. The
properties assumed in the analyses for the shear connectors are shown in Figure 6.16 and are
representative of load-slip data obtained from push-out tests on similar slab configuration. The
material model assumed for the steel beam was bilinear (linear up to the measured yield strength
of 410N/mm2 and then a plateau with strain hardening gradient of E/100). A trilinear model was
used for modelling the concrete which, however, does not capture the unloading behaviour
occuring at strains higher than 0.2% (which is the limiting strain to Eurocode 2).

85
Figure 6.16 Shear connector properties assumed in the model of the cellular beam

Nevertheless, the concrete strains were monitored to ensure this limiting strain was not exceeded.
Typical results for the analysis of the 15.3 m span cellular beam test are presented in Figure 6.17.
The load versus deflection and load versus end slip graphs obtained from the model are shown in
Figure 6.18 and are presented with the corresponding test results. As seen, a good agreement was
obtained.

Figure 6.17 Deformed shape (left) and strain in mid-span at the elongated cell (right) from the
analysis of the 15.3 m span cellular beam test using ANSYS

(a) Load versus deflection (b) Load versus end slip

Figure 6.18 Comparison between FE modelling and test results for 15.3 m span cellular beam
using ANSYS

86
The strain obtained from the FE model for the flanges of the steel section are compared against the
test strains in Figure 6.19. The overall agreement with the FE model is good and it captures the
development of strain in the flanges, but slightly under-predicts the bottom flange strain at high
levels of loading.

Figure 6.19 Comparsion of strains in FE model and test results for 15.3 m span cellular beam

6.4.3 Parametric analysis of cellular composite beams using ABAQUS


A parametric study was made using ABAQUS of two cellular beam configurations as follows:
Floor configuration n°1 : CP 60 decking

- Slab depth : 120 mm


- Span length : 3 m
- Distributed loading applied to the beam : qULS = 39.2 kN/m
- 19x100mm studs with through-deck welding
Floor configuration n°2 : CF 80 decking

- Slab depth : 150 mm


- Span length : 4.5 m
- Distributed loading applied to the beam : qULS = 61.4 kN/m
- 22 × 125mm studs welded through holes in the decking

A total of 8 configurations were studied with 1 or 2 studs per rib and for S355 or S460 steel grades
(refer to Table 6.5). Then, a large number of additional simulations were run in order to determine
the minimum degree of shear connection of cellular and full web beams under both unpropped and
fully propped conditions.
The minimum degree of shear connection is determined as the value of the degree of shear
connection that leads to:
 A moment of 0.95 Mpl is reached in the FE model (where Mpl is the plastic bending resistance for
the particular degree of shear connection)
 Deformation of the shear connectors is limited to 6 mm.
Table 6.5 Summary of geometrical data for configurations in the parametric analysis

Case L (m) Top T Bottom T D H Config 1-2 Ainf /Asup


1 10 IPE 330 HE 300 A 280 422 1 2.28
2 12 IPE 360 HE 340 A 310 470 1 2.29
3 15 IPE 450 HE 450 A 350 537 1 2.27
4 15 HE 360 A HE 360 B 320 484 1 1.29
5 12 IPE 450 HE 450 A 400 609 2 2.27
6 15 IPE 500 HE 550 A 480 719 2 2.25
7 18 IPE 600 HE 800 A 640 988.8 2 2.01
8 15 HE 450 A HE 450 B 410 613 2 1.24

87
In addition, parametric analyses were performed to determine the minimum degree of shear
connection of both cellular and solid web beams. Configurations 9 to 16 were the same as
configurations 1 to 8 except that the web openings were removed to compare the behaviour of
cellular beams and solid web beams.

Figure 6.15 shows a typical case of a graph of the ratio of the bending resistance obtained from the
FEA to the plastic bending resistance versus the minimum degree of shear connection for 6mm end
slip for four cases:

 15 m span asymmetric beam of 537 mm depth using CP60


 Opening diameter  0.65 × beam depth
 Cellular beam - H=537mm un-propped beam(H = opening diameter)
 Cellular beam - H=537mm propped beam
 Solid web beam H=0 (full web section) unpropped beam
 Solid web beam H=0 propped beam
These results in Figure 6.20(a) show that the minimum degree of shear connection of an un-
propped cellular beam was about 0.5 (50%) and that of a propped solid web beam was about 0.75
(75%).

All cases are summarised in Figure 6.15(b) in comparison to the requirements of Eurocode 4 for
asymmetric beams.

1.05 1.2
Cellular beams - Unpropped

1 1

0.95
Ratio Mmax,6mm/Mpl,EC4 (-)

Minimal degree of shear connection (-)

0.8

0.9
0.6
Symmetric
0.85 Asymmetric Ab/At = 2.25
Asymmetric Ab/At = 3
95% Limit 0.4
EN 1994-1-1 Rule Asymmetric
0.8
Config 3 - H = 537 - D = 350 - Unpropped New Rule Asymmetric
FEM Results Asymmetric
Config 3 - H = 537 - D = 0 - Unpropped
0.2 EN 1994-1-1 Rule Symmetric
0.75
Config 3 - H = 537 - D = 350 - Propped New Rule Symmetric

Config 3 - H = 350 - D = 0 - Propped FEM Results Symmetric


0.7 0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 5 10 15 20 25 30
Degree of shear connexion (-) Length (m)

(a) Typical results for a 15m span cellular (b) Summary of all FE cases from 10-18m
beam span

Figure 6.20 Minimum degrees of shear connection for unpropped cellular beams

The results of the FE models for Configuration 1 are presented in Table 6.6. It was found that the
difference between the minimum degree of shear connection for un-propped and propped beams
was 15 to 20% for the same cases. It was found that the minimum degree of shear connection of
un-propped cellular beams is 40 to 70% lower than the current Eurocode 4 requirements.

The effect of asymmetry of the flange areas was also found to be important, and the benefit of the
openings reduced with higher asymmetry. This is probably due to the position of the plastic neutral
axis which is lower in the section for asymmetric beams.

88
Table 6.6 Minimum degree of shear connection for cellular beams in comparison to the
equivalent solid web beam based on FEA

Span : Propped Beam Unpropped Beam


Beam span Flange EC4
Depth Solid Cellular Cellular
(m) Asymmetry Solid Web Minimum
Ratio Web Section Section

10 (Case 1) 24.0 2.28 0.65 0.55 0.52 0.45 0.74

12 (Case 2) 25.5 2.29 0.72 0.61 0.54 0.41 0.79

15 (Case 4) 30.9 1.29 0.72 0.38 0.54 0 0.72

15 (Case 3) 27.9 2.27 0.75 0.70 0.61 0.50 0.84

(a) 3 m span slab with CP60 decking with 120 mm deep slab

Span : Propped Beam Unpropped Beam


Beam span Flange EC4
Depth
(m) Asymmetry Solid Cellular
Solid Web
Cellular Minimum
Ratio Web Section Section

12 (Case 5) 19.7 2.27 0.73 0.48 0.52 0.37 0.78

15 (Case 8) 24.4 1.24 0.61 0 0.43 0 0.72

15 (Case 6) 20.8 2.25 0.81 0.63 0.67 0.42 0.84

18 (Case 7) 18.2 2.01 0.88 0.60 0.75 0.45 0.88

(b) 4.5m span slab with CF80 decking with 150 mm deep slab

Based on the results of this parametric study, it is propped that the minimum degree of shear
connection of a cellular beam may be related to that of the equivalent solid web beam according to
the formula:

η

= 1 
0.7ho Aft  η (6.1)
min, cellular  min, EC4
 h Afb 
where:

ho/h is the ratio of the opening diameter to beam depth

Afb/Aft is the ratio of the bottom to top flange areas


η min, EC4 is the minimum degree of shear connection for the equivalent solid web beam.

This equation shows that the effect of asymmetry of flange areas is to reduce the benefit of the
cellular cross-section. Also the formula does not apply outside a sensible limit of application which
is taken as: 0.8 ≥ ho / h ≥ 0.5. A further modification factor may be introduced for un-propped
beams -see WP6.

6.4.4 Parametric analysis of cellular composite beams using ANSYS


A second series of parameters was investigated using ANSYS: beam span, solid web beams or
beams with regular circular openings, propped or un-propped construction, asymmetry in the steel
section. Comparisons were made to the current Eurocode 4 shear connection requirements for
symmetric and asymmetric beams.

The symmetric cases were the basis for all the ANSYS analyses. Three spans were considered; 9
m, 12 m and 15 m. The steel section sizes chosen were IPE400, IPE500 and IPE600, respectively,
which have a span to depth ratio of 22.5 to 25. The slab was chosen as 150 mm deep with CF80
decking. The steel grade was S355 and the assumed concrete compressive strength was
30 N/mm2. The effective width of the concrete flange was taken as the beam span/4.

For the cellular beam cases, the diameter of the opening, h o ,were taken as 250 mm, 325 mm and
375 mm depending on the steel section size (ho/h between 0.67 and 0.69). The spacing of the
openings was constant at 1.6ho. The details of the asymmetric cellular beams are summarised in
Table 6.7. For the asymmetric solid web cases, the ratio of flange areas is given in Table 6.7.

89
Table 6.7 Asymmetric cellular composite beam cases considered

Span. Top tee Bottom tee Flange Diameter Opening Span/Depth


L [m] cut from cut from area ratio, of opening centre of beam
Afb/Aft ho [mm] spacing
[mm]
9 IPE400 IPE550 1.5 250 400 22.5
12 IPE500 IPE750x161 1.6 325 520 24

15 IPE600 IPE750x185 1.5 375 600 25

The results from the ANSYS analyses demonstrate that, for the parameters considered, the
proposed modifications in equation (8.1) to the Eurocode 4 degree of shear connection formulae for
cellular and un-propped beams provide very good accuracy, and in almost every case they are
conservative. This is clearly shown in Figure 6.21.

1
Degree of shear connection

0.8

0.6

0.4
EN1994-1-1 - Afb/Aft=1.5 approx.
0.2 ANSYS FE cellular - Afb/Aft=1.5 approx.
Modified EN1994-1-1 - Equation (8.2)
0
6 8 10 12 14 16 18
Span (m)

(a) Unpropped cellular beams

1
Degree of shear connection

0.8

0.6

0.4

EN1994-1-1 - Afb/Aft=1.5 approx.


0.2
ANSYS FE cellular - Afb/Aft=1.5 approx.
Modified EN1994-1-1 - Equation (8.1)
0
6 8 10 12 14 16 18
Span (m)

(b) Propped cellular beams

Figure 6.21 Comparison between current rules, proposed modification to the rules and ANSYS FE
results for asymmetric cellular beams of 9 m, 12 m and 15 m span

6.4.5 Additional deflections due to circular openings in cellular beams


Table 6.8 presents the results obtained for finite element analysis of the same cellular beams of 10
to 18m span with shear connectors corresponding to the minimum for the degree of shear
connection at the given span. It is based on un-propped asymmetric beams. The results are
compared to the deflection at 50% of the failure load, corresponding to a deflection between
span/300 and span/400. The beam span to depth ratio is in the practical range of 18 to 31 and so
is dominated by the flexural behaviour of the cellular beams.

90
The FEA results give an increase in deflection due to the openings in the range of 9 to 18%,
although this includes some effect of the difference of partial shear connection between the solid
web and cellular beam. The results are compared to a modified equation for the additional
deflection wadd relative to that of the solid web beam given by:

3
wadd h  h
 0.5 no  o    for s ≥ 1.4 ho (6.2)
wb  h  L

Where ho/h is the proportionate opening depth

L/h is the span: depth ratio of the beam

no is the number of circular openings in the beam span.

wb is the deflection of the solid web beam for the same loading

s is the centre-centre spacing of the openings.

It is shown that this new equation predicts an additional deflection of 12 to 17%, which agrees
reasonably well with the FEA results.

Table 6.8 Comparison of deflection of composite solid web beams and cellular beams and
comparison with approximate equation
Deflection at 0.5 % Increase in
Beam Span : Number
Flange ho Pu from FEA Deflection
span Depth of
Asymmetry h Solid Cellular
(m) Ratio Openings FEA Formula
Beam Beam

10 23.7 2.28 0.66 27 21.6 24.9 15.2% 16.6%

12 25.5 2.29 0.66 27 26.4 30.1 14.0% 15.2%

15 27.9 2.27 0.65 27 34.0 38.8 14.0% 13.4%

15 30.9 1.29 0.66 33 40.8 44.6 11.3% 15.4%

Data for beams with120mm deep slab using CP60 decking and 3m slab width

6.5 Concluding remarks for WP5.3


The following conclusions are made from the FEA regarding the minimum degree of shear
connection of cellular beams:

 For propped solid web beams, the current Eurocode 4 rules are reasonably accurate and no
changes are required.
 For unpropped beams, a factor of x 0.85 may be applied to the minimum degree of shear
connection for propped beams to the current Eurocode 4 rules



For cellular beams, a further reduction factor of x 1 
0.7ho Aft  may be applied.

 h Afb 
Example for a flange asymmetry of 2.26:1 and ho/h = 0.65, the minimum degree of shear
connection for a cellular beam = 0.8 x Minimum degree of shear connection for a solid web beam
to Eurocode 4.

6.6 WP5.5 Simplified finite element models


The simplified model considers the shear connectors as non-linear springs to connect the slab and
the steel beam. Figure 6.22(a) shows the model using 1D elements for the slab and steel section.
At the position of the shear connectors, the slab and the steel section are fixed to nodes at the
shear interface. The resulting pair of nodes is connected by non-linear springs to represent the
shear connector behaviour. Figure 6.22(b) shows this principle applied to a model where 2D
elements are used for the slab and the steel section. The following investigations were conducted
using the finite element software SOFiSTiK.

91
(a) Model using 1 D elements (b) Model using 2D elements

Figure 6.22 Simplified model using 1D and 2D elements for the slab and steel section

6.6.1 Comparison with beam tests in WP2


The analysis method was verified against Beam 2-10 with pairs of shear connectors in WP2. This
was a 6m span IPE 360 beam with CF 80 decking and pairs of shear connectors per deck rib. In the
beam tests, the strains in the steel section were measured every 900mm. Therefore, each interval
allowed the shear force introduced by the shear connectors in three ribs to be determined. The
obtained load-deflection curves compared to the test are shown in Figure 6.23. Similar behaviour
was obtained for Beam 2-09 with single shear connectors.

(a) Shear force per stud determined as function (b) Comparison of load-deflection in
of end slip test and FEA
Figure 6.23 Comparison of simplified FEA model versus test beam 2-10 of WP2

6.6.2 Parametric study


Having established that the simplified FE model can be used accurately, a parametric study was
made of different sizes and spans of beams as shown in Figure 6.24. A simplified behaviour of the
shear connectors was defined based on the push tests in WP1.

(a) Beam sizes in parametric study (b) Comparison of load-deflection in test and FEA
Figure 6.24 Beam sizes used in the study and idealised load-slip behaviour of the shear
connectors

The analysis using CF80 decking showed that the average force of shear connectors that had
developed plasticity was about 92% of the failure load of the rib, Pe1 when the beam reached its
maximum load. Therefore, the simplified load-slip curve used a shear resistance of 0.92 Pe1. As
shown in Figure 6.25, the simplified load-slip behaviour lead to almost identical results as the more
accurate behaviour. The obtained results for the slip were very accurate until a slip of 5mm was
92
exceeded. Similar analyses were performed for CP60 decking, which are reported in the Deliverable
to WP5.5.

(a) Comparison of moment-deflection (b) Comparison of slip-deflection


Figure 6.25 Results of simplified FE analysis of 6 to 12m span composite beams

6.7 Concluding remarks for WP 5


A wide range of finite element models were created to analyse the tests in WP2 and 3 and to carry
out parametric studies using idealised load-slip characteristics of the shear connectors. These
were:

 ABAQUS models that either used solid elements for the shear connectors and beams or shell
elements with springs to represent the shear connector behaviour. It was found that models
including the decking were unstable and in later models, the decking was removed with no loss
of accuracy.
 ANSYS models with shell elements to model the steel profile and springs to model the shear
connectors. However, ANSYS cannot model the declining concrete strength at high strain.
 SOFiSTiK models using 1-D and 2-D elements and springs to represent the shear connectors.
This proved to be an accurate ‘tool’ and was used in the parametric studies for different beam
sizes and spans.
ABAQUS was also used to model cellular beams with variable degree of shear connection so that
the bending resistance at a limiting slip of 6mm could be determined. This leads to a reduced
minimum degree of shear connection of cellular beams in comparison to the equivalent solid web
beams.

6.7.1 Exploitation and impact of the research results


The finite element models proved to be accurate in the way they modelled the behaviour of the
test beams and were used in the parametric studies in WP5 and 6.

93
94
7 WP6: Improved shear connection rules for composite beams

7.1 Objectives of WP6


The objective of this Work Package was to develop improved shear connection rules in comparison
to the existing rules in EN 1994-1-1, for the following cases:

 Un-propped beams taking account of the self-weight of the beam and slab acting on the steel
beam.
 Semi-ductile shear connectors with lower slip (3 mm typically) and also super-ductile shear
connectors with higher slip (taken as 10 mm) at their design resistance.
 Rules for shear connector performances as a function of end slip.
 Beams designed for less than their plastic bending resistance (i.e. for lower utilisation at the
ultimate limit state where the design is strongly influenced by serviceability conditions).
 Cellular beams with large web openings where the web does not contribute significantly to the
bending resistance, and so different shear connection rules may apply.
 Primary beams, as influenced by the effective transverse reinforcement.
 Edge beams with and without end anchorage of the reinforcement.
These rules were based on the results of the numerical models established in WP 5. The existing
shear connection limits in EN 1994-1-1 are illustrated in Figure 7.1.

Figure 7.1 Minimum shear connection rules in EN 1994-1-1 for symmetric and asymmetric
beams and S275 and S355 steel grades

7.2 Comparison of initially planned activities and work accomplished


This WP used the results obtained from the testing and finite element modelling activities in the
earlier WPs. To avoid making too many fundamental changes to Eurocode 4, it was decided that
one approach is to apply modification factors to the existing clauses. For example, for the minimum
degree of shear connection in clause 6.6.1.2, the minimum degree of shear connection includes a
multiple for the following:
 Unpropped beams.
 Utilisation factor in bending at the ultimate limit state.
 Limiting end slip ranging from low to high deformation capacity.
 Opening diameter of cellular beams.

95
7.3 Description and impact of the research results- WP6.1

7.3.1 Elastic design of composite beams


Elastic design is used to check deflections and stresses of composite beams at the serviceability
limit, but may also be used at the ultimate limit state when the cross-section does not meet the
Class 1 or 2 criteria in EN 1993-1-1: Eurocode 3. Elastic design is also used for composite beams
with non-ductile shear connectors that fail to meet the 6 mm limiting slip in Eurocode 4.

Slip at the interface between the beam and slab leads to deformations along the beam, as shown in
Figure 7.2. Furthermore, the lower the shear connector stiffness, the higher the deflection due to
slip, but the lower the shear forces experienced in the outer shear connectors.

The first theory for the effects of partial shear connection on the design of composite beams with
partial shear connection was developed by Newmark, who presented a solution of a differential
equation linking slip and deflection for beams. The differential equation may also be solved for the
general load case using finite difference method and many papers have presented this approach,
such as by Ranzi..

The following theory shows how the stiffness of the shear connectors may be included in the
calculation of the effective stiffness of composite beams and also in the longitudinal shear forces. It
is based on the theory presented by Lam, Elliott and Nethercot (Proc ICE,2000), which was
developed for composite beams supporting precast concrete slabs. The theory may be explained by
making the assumption that the interface slip follows a simple cosine function along the beam to a
maximum slip at the ends of the span. The actual slip distribution along the beam will differ with
applied loading but comparison with full-scale beam tests shows that these simplifications are
reasonable for deflection calculations.

F
Mc ssc
s-ds

s F+dF hc
M hsc
Mc+dMc hp
Ms
F

F+dF
I section Ms+dMs M+dM
h

x dx

(a) Side view of composite beam showing slip of shear connections (b) Cross-section through composite beam

Figure 7.2 Forces and displacements in a composite beam as affected by slip

Based on this theory, the effective inertia of the composite beam as a function of the shear
connector stiffness is:

Ic
ys  0.5 hc  hp 2  Anc 
I I    
(7.1)
eff s n  A 2  E  A 
1  c   π   s  c 
 nA  L   k / s  n 
 s  sc  
The end slip may be presented as a function of the inertias of the composite section, Icomp, for rigid
shear connectors, and of the steel section, Is, making the assumption that the term Ic/n is small
and can be neglected. The end slip is given as follows for an applied moment, M:

96
L /   ys  0.5 hc  h p  M
s (7.2)
 I  L   k / ssc   Ac  n As 
2
EI comp  s    . . A A 
 I comp     E s  s c  

Aa is the cross-sectional area of the steel beam

Ac is the cross-sectional area of the concrete slab.

Es is the elastic modulus of steel

Ec is the elastic modulus of concrete.

hc is the depth of the concrete over the deck profile

hp is the depth of the deck profile

Is is the second moment of area of the steel beam

Ic is the second moment of area of the concrete slab

Icomp is the second moment of area of the composite section for rigid shear connectors (k= ∞)

k is the stiffness of the shear connectors (in kN/mm -see below)

L is the beam span

n Es/ Ec

ssc is the spacing of the shear connectors along the beam

ys is the elastic neutral axis depth of the steel section from the top of the section

7.3.2 Study on shear connector stiffness


The initial elastic stiffness and secant stiffness of single and pairs of shear connectors was obtained
from the results of push-out tests in this project. Representative tests from the series were
selected for this study all using C30/37 concrete; 19 mm diameter and 125 mm high shear
connectors in 80 mm deep trapezoidal decking, and 19 mm diameter and 100 mm high shear
connectors in 58 mm deep trapezoidal decking and 56 mm deep re-entrant decking.

The secant stiffness is obtained at the maximum value of resistance (defined as the first peak in
the load-slip plot), while the initial elastic stiffness is obtained at 60% of this value. This is an
appropriate serviceability condition for the elastic forces in the shear connectors for unpropped
beams. The shear connector stiffness obtained from push tests are presented in Table 7.1. From
these tests, the representative stiffness may be taken as 70 kN/mm for single shear connectors
and 100 kN/mm for pairs of shear connectors per deck rib. For low load levels, the elastic stiffness
for single shear connectors will be higher than these values.

Table 7.1 Summary of shear connector stiffness from push tests (values rounded)

Decking Configuration Shear No. of Initial Secant


connectors studs elastic stiffness
per rib stiffness [kN/mm]
[kN/mm]
80mm deep decking with wide rib 19mm dia. x 1 70 35
125mm high
2 70 55
58mm deep decking with narrow rib 19mm dia. x 1 70 25
100mm high
2 110 30
56mm deep decking with wide re-entrant 19mm dia. x 1 60 30
rib 100mm high
2 120 50

7.3.3 Elastic stiffness of long span beams in WP3


The effective inertia for the 11.2 m span asymmetric beam Ieff is 74% of the composite inertia,
Icomp for rigid shear connectors For an imposed load of 5 kN/m2, the end slip was obtained as
0.53 mm from equation (9.2). Using this stiffness, the shear connectors would reach their elastic
97
limit at a load of about 10 kN/m2, which agrees well with the test and is equivalent to 47% of the
test failure load.

The calculated deflection of the fully composite beam with rigid shear connectors is 12 mm and the
deflection of the composite beam with flexible shear connectors (using k= 70 kN/mm) is 16.2 mm.
Therefore, the additional deflection due to end slip is 35% of the deflection of the composite beam
with rigid shear connectors. The test deflection was approximately 16 mm at this load, which is in
good agreement with the theoretical deflection using this flexibility as seen in Figure 7.3(a).

A second test was performed on a cellular beam of 15.3 m span with the same slab depth, deck
profile and shear connectors and loading system as the previous test. The calculated degree of
shear connection was 36%. The third cycle of loading to 12 kN/m2 is presented in Figure 7.3(b).
The behaviour was elastic up to a load of about 7.5 kN/m2 and the end slip became irreversible at a
load of about 10 kN/m2, which is equivalent to 49% of the failure load.

(a) 11m span asymmetric beam (b) 15m span cellular beam

Figure 7.3 Load versus deflection for long span beams up to twice working load and comparison
with the elastic theory

7.3.4 Parametric study on end slip at serviceability


A series of finite element analyses (FEA) was carried out for IPE 270 to IPE 600 beams using the
generalised finite element program ANSYS, in which the shear connector load-displacement curve
was input for single shear connectors at 300 mm spacing. The beam data are presented in
Table 7.2 and the FE results are presented in Table 9.3, in which Mu is the plastic bending
resistance obtained from the FE model and Pu is the equivalent line load at failure.

An imposed load of 0.3 Pu is taken as being typical of the service load for unpropped composite
beams, and a load of 0.5 Pu is taken as being typical of the service load for propped composite
beams. It is shown that for unpropped beams, the end slip varies only slightly with beam span and
does not exceed 0.7 mm, which is within the elastic range. For propped beams, the end slip
increased from 1.1 to 1.5 mm for 9 to 15 m span, which is the post–elastic range and would lead
to higher deflections..

The FEA results in Table 7.1 may be compared to the theory in this paper for the two load levels
and the results are shown in Table 7.2. It is shown that the comparisons for un-propped beams
and for shorter span propped beams are good. However, the non-linear behaviour of the shear
connectors is apparent for propped beams with spans of more than 10 m.

Comparison is also made with the former BS 5950-3, which gave a formula for the additional
deflection due to partial shear connection, whereas no equivalent formula is given in Eurocode 4.
For both unpropped and propped beams, the deflection according to BS 5950-3 is 10 to 25%
higher than given by the theory in this project.

98
Table 7.2 Data of the composite beams used in the FE models

Beam Parameter

Beam Span Beam Opening Degree of Slab width Applied


Size diameter shear moment/Bending
(S355) connection resistance
9m IPE 400 300 mm 35% 2.25 m 0.26
12 m IPE 500 350 mm 34% 3m 0.37
15 m IPE 600 400 mm 33% 3.75 m 0.47

Loading = 6 kN/m2
Shear connectors at 300 mm spacing
Slab depth =150 mm
Profile depth = 80 mm

Table 7.3 Deflection of composite beams at 30% and 50% of their maximum load, Pu
corresponding to the bending resistance Mu based on finite element results

Load of 0.3 Pu Load of 0.5 Pu


Pu
Span [m] Section Mu [kNm] [kN/m] δ [mm] s [mm] δ [mm] s [mm]
6 IPE270 308.9 68.6 11.4 0.53 19.0 0.88
9 IPE400 712.6 70.4 17.4 0.60 29.5 1.12
12 IPE500 1168.5 64.9 24.5 0.63 41.7 1.28
15 IPE600 1803.1 64.1 32.3 0.66 55.5 1.48

Table 7.4 Comparison of deflections and the elastic theory in the FE models

Theory in this Deliverable Comparison with BS 5950-3


Load of 0.3 Pu Load of 0.5 Pu = 0.3 Pu = 0.5 Pu
Span
[m] Section δ [mm] s [mm] δ [mm] s [mm] δ [mm] δ [mm]
6 IPE270 10.1 0.55 16.8 0.91 12.7 21.2
9 IPE400 16.0 0.61 26.7 1.02 17.8 29.6
12 IPE500 23.1 0.64 38.4 1.06 25.8 43.0
15 IPE600 30.5 0.67 50.8 1.12 33.6 56.0

The same finite element model was used for cellular beams of 9 to 15 m span with opening depths
of 67% to 75% of the depth of the beam. The models were analysed for un-propped beams with
rigid and flexible shear connectors in order to determine the effect of partial shear connection on
deflections. The degree of shear connection was about 35% in all the models and ratio of the
moment due to the imposed load to the bending resistance was in the range of 27% to 47% for the
9 to 15 m spans. The same models were repeated for solid web beams in order to determine the
effect of the openings for both rigid and flexible shear connectors.

The results were compared to the theoretical flexural model that was proposed with an additional
small term due to the additional shear deflection caused by the openings. Although this term is
small, it is nevertheless important when comparing the differences between the results. It is shown
that the theoretical method with this adjustment for shear displacement agrees well with the FE
results, as already observed in the cellular beam test.

99
Table 7.5 Comparison of deflections of solid web beams and cellular beams from FEA, theory
for rigid and flexible shear connectors

Solid web beam - k = 1000 kN/mm Cellular beam - k = 1000 kN/mm Add.
defln. due
Beam FEA Theory - Additional Total FEA Theory Additional Total
to
Span result rigid Shear deflection result rigid shear defln.
openings
(m) shear deflection from shear deflection
based on
connectors theory connectors
FEA, wadd
9 8.9 7.9 0.5 8.4 11.1 9.0 1.0 10.0 2.1 (23%)
12 18.9 17.4 0.8 18.2 22.8 19.7 1.6 21.3 3.9 (20%)
15 32.5 30.0 1.1 31.1 38.6 33.9 2.0 35.9 6.1 (19%)
(a) Effectively rigid shear connectors

Solid web beam - k=70kN/mm Cellular beam - k=70kN/mm Add. defln.


Beam FEA Theory Additional Total due to
FEA Theory Additional Total
Span result including Shear openings
defln. result inc. shear defln.
(m) based on
flexibility deflection flexibility deflection
FEA, wadd
of SC of SC
9m 11.1 10.1 0.5 10.6 13.1 11.2 1.0 12.2 2.0 (18%)
12m 22.5 21.0 0.8 21.8 26.2 23.3 .6 24.9 3.7 (16%)
15m 37.5 35.1 1.1 36.2 43.2 39.0 2.0 41.0 5.7 (15%)
Difference due to flexibility of shear connectors
9m 2.2 2.2 - 2.2 2.0 2.2 2.2
12m 3.6 3.6 3.6 3.4 3.6 - 3.6
15m 5.0 5.1 5.1 4.6 5.1 - 5.1
(b) Flexible shear connectors

7.3.5 Sensitivity study on minimum degree of shear connection at the


serviceability limit state
The finite element analyses were extended to consider at what point a limiting slip of 1mm is
exceeded for beams of 6 m to 15 m span with a span: depth ratio in the range of 22 to 25. An
initial analysis was performed where to determine the working loads for these beams based on
their plastic bending resistance for the particular degree of shear connection. The analyses are
repeated by adjusting the degree of shear connection so that the end slip did not exceed 1 mm.

The serviceability load in this analysis was first determined from a moment of: (Mpl –1.35Msw)/1.5
for un-propped beams and Mpl/1.5 for propped beams, where Mpl is the plastic resistance of the
composite beam for partial shear connection and Msw is the moment due to the self-weight of the
beam and slab.

It is argued that the statistical likelihood of repeated loading of the magnitude of the full
serviceability load is low and therefore the cumulative effects of slip will not be significant.
Therefore, the load at which the slip is calculated is taken as: 0.8 x (Mpl –1.35 Msw)/1.5 for
unpropped beams and 0.8 Mpl/1.5 for propped beams.

7.3.6 Results for symmetric beams


For symmetric beams, the results for the minimum degree of shear connection for a 1 mm slip limit
at the serviceability limit state are presented in Table 7.6. These minimum values were found to be
in the range of 0.38 to 0.47 for propped beams and 0.26 to 0.3 for unpropped beams (i.e. about
20% less than for propped beams).

100
Table 7.6 Minimum degree of shear connection for symmetric beams based on 1mm slip limit
at the serviceability limit state

Span [m] Beam (in Number of Unpropped beams Propped beams


S355 shear
Minimum degree of shear connection for 1 mm slip
steel) connectors
calculated for a serviceability moment of:
per deck
rib 0.8(Mpl –1.35 (Mpl 1.35 0.8 Mpl/1.5 Mpl/1.5
Msw)/1.5 Msw)/1.5
6 IPE270 1 0.26 0.37 0.29 0.40
2 0.32 0.43 0.37 0.48
9 IPE400 1 0.27 0.37 0.34 0.46
2 0.32 0.46 0.39 0.53
12 IPE500 1 0.26 0.37 0.37 0.48
2 0.32 0.43 0.43 0.56
15 IPE600 1 0.23 0.35 0.38 0.50
2 0.30 0.43 0.47 0.55
Shear connector stiffness of 70 kN/mm for single shear connectors and 100 kN/mm for pairs of
shear connectors
Mpl is the plastic resistance of the beam for partial shear connection
and Msw is the moment due to the self-weight of the beam and slab

Based on these results for a maximum serviceability slip of 1 mm, it is proposed that the cut-off in
the minimum degree of shear connection for symmetric composite beams could be taken as 0.3 for
unpropped beams and 0.4 for propped beams. The 0.4 minimum limit is the same as for the
existing shear connection rules in Eurocode 4, and so a relaxation in the minimum degree of shear
connection is made only for unpropped beams.

7.3.7 Results for asymmetric beams


Beams with 3:1 asymmetry in flange areas were created by multiplying the top flange area by 0.5
and the bottom flange area by 1.5 to keep to the same area as the parent IPE beam. The same
approach was adopted but in this case the minimum degree of shear connection is higher than for
symmetric sections. Results of these analyses are presented in Table 7.7. For propped asymmetric
beams, the minimum degree of shear connection is in the range of 0.40 to 0.58 and an unpropped
beams is in the range of 0.37 to 0.46.

Table 7.7 Minimum degree of shear connection for asymmetric beams based on original IPE
section with flange areas of 3:1 based on 1 mm slip limit at the serviceability limit
state

Span Asymmetric Number of Unpropped beams Propped beams


[m] beam based shear
Minimum degree of shear connection for 1mm
on original connectors
slip calculated for a serviceability moment of:
section (in per deck
S355 steel) rib 0.8 (Mpl 1.35 (Mpl 1.35 – 0.8 Mpl/1.5 Mpl/1.5
Msw)/1.5 Msw)/1.5
6 IPE270 1 0.37 0.48 0.40 0.51
with Afb/Aft =3 2 0.43 0.53 0.48 0.59
9 IPE400 1 0.39 0.54 0.46 0.61
with Afb/Aft =3 2 0.46 0.60 0.53 0.68
12 IPE500 1 0.38 0.51 0.48 0.60
with Afb/Aft =3 2 0.43 0.59 0.53 0.69
15 IPE600 1 0.38 0.51 0.50 0.64
with Afb/Aft =3 2 0.43 0.58 0.58 0.73

101
As a good approximation for unpropped beams, the minimum limit on the degree of shear
connection to control slip at the serviceability limit state may be taken as:

Min η = 0.2 + 0.1(Afb/Aft) > 0.3 (7.3)

where: Afb/Aft is the ratio of flange areas

Similarly for propped beams, the minimum limit on the degree of shear connection to control slip at
the serviceability limit state may be taken as:

Min η = 0.3 + 0.1(Afb/Aft)) > 0.4 (7.4)

These cut-off values should be applied to the general rules for the minimum degree of shear
connection at the ultimate limit state irrespective of the limiting end slip.

7.4 WP6.2 Rules for partial utilisation

7.4.1 Current shear connection rules based on full utilisation in bending


The minimum degree of shear connection requirements in EN 1994-1-1 for composite beams with
symmetric steel sections is given by the following relationships:

For Le < 25m:

355
  1 0.75  0.03Le 
fy
(7.5)

and 1.0 ≥ η ≥ 0.4

where η is the degree of shear connection, fy is the yield strength of the steel section and Le is the
distance between points of zero bending moment.

For certain steel decking and shear connector configurations (deck heights up to 60 mm, one shear
connector per deck rib, decks spanning transverse to the beam and continuous across it),
alternative requirements that consider more ductile behaviour are allowed (cl. 6.6.1.2-3):

355
  1 1.0  0.04Le 
fy
(7.6)

Again 1.0 ≥ η ≥ 0.4

In one of the two studies for the calibration of EN 1994-1-1 by Aribert, it was suggested that a
lower limit would be possible, but this was not adopted in the rules.

The degree of shear connection in composite beams is calculated from:

Rq

min Rs , Rc  (7.7)

where Rq is the force transferred by the shear connectors located between mid-span and support in
uniformly loaded and simply supported beams, Rs is the axial resistance of the steel section and Rc
is the compression resistance of concrete. The reduced demand for slip with lower steel grades
suggested by the current formulae is therefore particularly relevant to cases where Rs is greater
than Rc and the steel section is the dominant term.

7.4.2 Shear connection requirements based on partial utilisation of composite


beams in bending
The objective of this study was to define and justify less restrictive shear connection requirements
for composite beams based on their partial utilisation in bending. In order to achieve this, a
parametric finite element study was carried out in ANSYS. Beam elements were used to represent
both the slab (using the depth above the profile only) and the steel beam (using a shape sub-set to
input different beam sizes). Rigid elements were used to connect the beam and slab elements at
102
the shear connector positions, and these were broken at the interface between the beam and slab
with a spring that was constrained to move only parallel to the beam, as shown. This spring initially
had zero length.

The slab width was taken as the effective breadth of the slab (the minimum of span/4 or the design
beam spacing, which was set to 3 m for this exercise). A bi-linear material model was used for
steel. No allowance was made for strain hardening, since it is not expected to have any influence
on the results of this study.

A bi-linear model was also used for the concrete with an initial gradient of the modulus of elasticity
(taken as 21 kN/mm² to reflect combined short- and long-term loading) and a plateau in
compression at 0.85 fck /c (where fck is the characteristic cylinder strength, taken as 25 N/mm²).
For concrete in tension, where cracking would normally occur, a bi-linear relationship was again
used, with the same initial gradient but a plateau at 10% of the compressive plateau.

The force-deflection behaviour of the shear connector spring was defined using either a bi-linear or
a multi-linear model. The model was loaded by applying a uniformly distributed load to the top
surface of the concrete slab. Failure is defined as the point when the beam achieves 95% of its
plastic moment capacity. This typically occurs at deflections in the region of span/100, and this
definition of failure is supported by the work of Aribert.

Typical strains at this level of load are between 2.4 and 3.6 times the yield strain (of 355 N/mm²
steel) and a strain of 0.75  10-3 in the concrete (the maximum strain allowed in the concrete is c,3
= 1.75  10-3). These strain levels imply that neither strain hardening of the steel nor concrete
crushing has to be included in the analysis. The model was validated against the 11.2m span
fabricated beam test in WP3 and the comparison is presented in Figure 7.4.

20
18
16
14
Load (kN/m2)

12
10
8
6
4
2
0
0 50 100 150 200
Deflection (mm)

Test FE FE - 20 MPa concrete strength FE - k=100 kN/mm

Figure 7.4 Load versus deflection curves obtained from FEA and tests on the 11.2m span
fabricated beam in WP3

7.4.3 Parameters investigated


Uniformly loaded composite beams of spans from 9 m to 18 m were considered in this study. Only
propped construction was considered since this would give more conservative results in terms of
the force transferred to the shear connectors. The span to depth ratio of the steel sections
considered was between 22 and 25 approximately. The steel grade was taken as S355.

The cases were analysed for both single and pairs of shear connectors per deck rib at 300 mm
spacing. The slip capacity of the shear connectors was assumed to be 6 mm for a single shear
connector (ductile behaviour) and 3 mm for a pair of connectors (semi-ductile behaviour), based
on the observation that use of pairs of shear connectors might reduce the deformation capacity
(Mottram and Johnson). Although the semi-ductile behaviour is unrealistic based on recent tests by
the SCI and the tests carried out within WP1, a broader range of behaviours with regard to
deformation capacity was investigated. For the 15 m to 20 m span cases, the analysis was
repeated for single shear connectors with 10 mm slip capacity in order to demonstrate the effect of
using shear connectors with higher deformation capacity.
103
The results from the parametric study are presented in Figure 7.5 for symmetric sections. The
horizontal axis is the utilisation ratio (UR) at the ultimate limit state, which is the ratio of the
applied moment to the plastic bending resistance of the composite beam for the particular degree
of shear connection. The vertical axis is the multiplier that may be applied to the shear connector
resistance to obtain a plastic bending resistance equal to the applied moment.

This graph shows the multiplier reduces with the span of the beam. For beams of 18 m span, it is
shown that a multiplier equal to UR2 is a conservative estimate of the reduction in the minimum
degree of shear connection.

1
Shear connection multiplier

0.8

0.6

0.4

0.2

0
0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
Utilisation Ratio

9m span 12m span 15m span 18m span (UR)^2

Figure 7.5 Parametric study results for single shear connector per deck rib and 6 mm slip
capacity (UR = Utilisation ratio = M/Mpl)

Therefore, at lower utilisations in bending, a reduction in the shear connection required at full
utilisation is possible. This could be in the context of the following relationships that sue the
existing Eurocode 4 formulae.

For shear connectors within ribs of deck profiles, where a 6 mm slip capacity can be demonstrated
(ductile behaviour), it is proposed that the minimum degree of shear connection is given by:
2
M
Le ≤ 20 m:    Ed  1  355  0.75  0.03 L e  
  
(7.8)
M 
 Rd   fy 

but   0.4  M Ed 
 (7.9)
M
 Rd 

7.4.4 Sensitivity of results and current limitations


The current study was based on the assumption that single shear connectors in the deck ribs can
achieve 6 mm slip, while pairs of connectors can reach 3 mm slip. This is a generalised assumption
made for the purposes of this study. The tests in this project show that the characteristic slip
capacity of headed stud shear connectors singly placed in the ribs of steel decks can be
significantly greater than 6 mm.

Finally the effect of taller shear connectors used ina deeper profile and a thicker slab is shown in
Figure 7.6. In this case, it is shown that for 19 mm diameter  120 mm high shear connectors in
80 mm deep decking, slightly more onerous results would have been obtained. However, the
proposed relaxation to the current rules is still valid.

104
1

Shear connection multiplier


0.8
0.6
0.4
0.2
0
0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
Utilisation Ratio

60mm deck - 130mm slab - 19x95 stud


80mm deck - 150mm slab - 19x120 stud

Figure 7.6 Parametric study results showing the effect of different deck/slab and shear
connector dimensions (12 m span case)

Special considerations need to be made for beams loaded by point loads and especially off-centre
loads, where the critical section is not at mid-span.

7.4.5 Role of asymmetry in the flange areas of the steel section


For asymmetric steel sections where the area of the bottom flange is significantly greater than the
top flange, the same concept applies. This is demonstrated in Figure 7.7 for an assumed slip
capacity of the shear connectors equal to 6 mm and a bottom flange area three times that of the
top flange. However, the reduction factor (MEd/MRd)2 proposed for symmetric beams is slightly un-
conservative (by about 5%) for very long span (18m span) and highly asymmetric beams.

1
Shear connection multiplier

0.8

0.6

0.4

0.2

0
0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
Utilisation Ratio

12m span 18m span (UR)^2

Figure 7.7 Parametric study results showing the effect of beam asymmetry (Afb/Aft=3)

7.4.6 University of Luxembourg results for utilisation end slip


The concept of lower degree of shear connection for beams that are not fully utilised in bending
was also investigated by the University of Luxembourg. The results are presented for a 12 m span
beam using an IPE 450 section (in S355 steel) which were analysed using the methods described in
WP5.5.

In Figure 7.8, the FE analyses of the beam were performed for beams with different degrees of
shear connection and utilisation ratios of 0.5 to 0.95. The end slip varied between 0.5 and
12.5mm. For a limiting slip of 6mm, a horizontal line may be drawn which intersects the curves for
different utilisation ratios. It leads to the relationship in Figure 7.8, which shows that equation
(9.8) may be safely used.

105
Figure 7.8 Results from the University of Luxembourg in terms of utilisation factor and slip and
degree of shear connection

Figure 7.9 Graph showing the effect of utilisation factor on the minimum degree of shear
connection for 6mm slip

7.4.7 Proposed adaptation of Eurocode rules for partial shear connection


The current Eurocode 4 formulae for the minimum degree of shear connection in clause 6.6.1.2
may be considered to be sufficiently accurate for propped beams of symmetric and asymmetric
shapes in steel grades up to S355. The following modification factors are proposed for the
minimum degree of shear connection of unpropped composite beams and cellular beams, which
includes the following factors:

 Un-propped construction, expressed as a function of the proportion of loading acting on the


steel beam in construction in relation to the plastic bending resistance. Msw,Ed/Mpℓ,Rd, where
Msw,Ed is the moment due to self-weight loads and Mpℓ,Rd is the plastic bending resistance of the
composite beam for the degree of shear connection used.
 Utilisation of the composite beam in bending at the ultimate limit state based on the ratio: MEd/
Mpℓ,Rd,, in which MEd is the applied moment due to the self-weight of the beam and slab.
 Limiting slip capacity of the shear connectors, based on the end slip for development of the
design resistance of shear connectors (default value, Slim = 6 mm).

106
 Cellular beams with regular circular openings expressed in terms of the proportionate depth of
opening in a cellular beam, ho/h , in which the bending resistance is determined for the
cross-section at an opening. This is a function of the top and bottom flange areas, Afb/Aft.
A combined formula for minimum degree of shear connection, η, in the general case with shear
connectors satisfying the 6 mm slip criterion is:

2
 Msw, Ed   MEd   0.7ho Aft 
η = ηEC4. 1   .  . 1   < 1.0 (7.9)
 Mpl, Rd   Mpl   h Afb 

where ηEC4 is the minimum degree of shear connection obtained from Eurocode 4 clause 6.6.1.2,
which applies to propped beams with a limiting slip capacity of 6mm.

 Msw, Ed 
The effect of unpropped construction is given by the term: 1   which is in the range of
 Mpl , Rd 
0.75 to 0.85 and may be taken conservatively as a single value of 0.85. At the ultimate limit state,
the utilisation of the section in bending has an important effect of reducing end slips and a factor of
 
MEd / Mpl 2 allows for this. A limit of MEd / Mp ≥ 0.7 is used in the above equations.

Therefore for un-propped beams without web openings, the minimum degree of shear connection is
given by:
2
 MEd 
η = 0.85 ηEC4.   < 1.0 (7.10)
 Mpl 

For shear connectors with higher or lower end slips, Slim, the minimum degree of shear connection
is reduced and the factor 6 / Slim 
0.5
allows for this in the following formula:

0.5
 MEd   6 
η =0.85 ηEC4.   .   < 1.0 (7.11)
 Mpl   Slim 

For shear connectors satisfying a higher slip criterion, the utilisation of the section in bending has a
less important effect than for shear connectors satisfying the 6mm slip criterion and so a linear
factor of MEd / Mpl  is introduced to allow for this in equation (9.11). For a limiting 10mm end slip
in un-propped composite beams without web openings, the above equation becomes:

 MEd 
η = 0.66 ηEC4.   . < 1.0 (7.12)
 Mpl 

The effect of the openings in cellular beams is to further reduce the minimum degree of shear
connection in equations (9.10) to (9.12) according to the factor:. 1  0.7ho Aft  . No reduction is
 h Afb 

permitted for isolated or non-circular openings or for openings with diameter, ho< 0.5 h.

In using the above formulae, it is proposed that the current minimum cut-off limit of 0.4 for
propped beams is modified according to:


η> 0.4 1 
Msw, Ed   MEd   Afb 
 .   .  0.6  0.4  and η > 0.24 (7.13)
 Mpl , Rd   Mpl , Rd   Aft 

For un-propped symmetric beams, the cut-off is conservatively given by:

 MEd 
η = 0.34   and η > 0.24 (7.14)
 Mpl 

The cut-off limit is also required to ensure adequate serviceability performance of composite beams
which avoids the cumulative effect of slip on deflections under repeated loading. End slips reduce
linearly with utilisation factor in elastic design and so the factor MEd / Mpl, Rd is introduced. The
limit is 0.4 for propped symmetric beams, which is the same as Eurocode 4. This cut-off limit
increases with the asymmetry of the cross-section.

107
The following examples show the use of these formulae:

1. Example for long span un-propped asymmetric beam:

Span L= 15m, steel strength fy =355 N/mm2, flange area ratio Afb/Aft =3, end slip Slim=6mm:

ηEC4 = 1- (0.3 - 0.015 x15) = 0.92

 For un-propped beam, η = 0.92 x 0.85 = 0.79 (79%).


 The cut-off limit is: η = 0.4 x 0.85 x (0.6 + 0.4 x 3) = 0.61, which is less than 0.79, and so
the minimum degree of shear connection of 0.79 (79%) will control for this asymmetric beam.
 For partially utilised beam with MEd/Mpl,Rd = 0.8, η = 0.79 x 0.82 = 0.50 (50%)
 The cut-off limit is: η = 0.4 x 0.85 x 0.8 x (0.6+ 0.4 x 3) = 0.49, which is less than 0.5, and
so the minimum degree of shear connection of 0.5 (50%) will control for this load level.
2. Example for long span un-propped asymmetric cellular beam:
For data as above with ho/h = 0.7 and slim = 6mm:

 For unpropped beam, η = 0.79 x (1- 0.7x 0.7/ 3) = 0.66 (66%)


 The cut-off limit is: η = 0.61 as above, which is less than 0.66.
 For partially utilised beam with MEd/Mpl,Rd = 0.8, η = 0.66 x 0.82 = 0.42 .
 The cut-off limit is: η = 0.4 x 0.85 x 0.8 x (0.6 + 0.4 x 3) = 0.49, which is more than 0.42,
and so the cut-off limit of 0.49 (49%) will control for this partially utilised asymmetric cellular
beam.

7.5 Beams loaded by point loads


Composite secondary beams are uniformly loaded. However, in some cases, beams can be loaded
by a combination of uniform loading and point loads. In such cases, the effect of the point loads
cannot be easily assessed, because it depends on their position in the span and also their relative
magnitude. Primary beams are loaded by point loads predominantly. The results from FE analyses
on point-loaded beams are compared with uniformly loaded beams. The most common loading
arrangement is when the loads are positioned at a distance of one third the beam span from the
support.

The ANSYS FE model used for the purposes of D6.2 was used. Although this model was validated
for uniform loading, it is further validated here against test beam 2-09 of WP2 which were tested in
by two point loads, details of which are given in Figure 7.10.

Figure 7.10 Test setup for series 2-09 and 2-10 – side view

The 6 m span IPE360 beam combined with a 150 mm deep slab had a span to depth ratio of about
17. The values of the yield strength of the steel and the compressive strength of the concrete
(target grade) were taken as 375 N/mm2 and 48.6 N/mm2 respectively, based on measured
values. The shear connector resistance used in the FE analysis was 68 kN with an elastic stiffness
of 70 kN/mm. The number of shear connectors between the critical section at the position of the
load and the support was 8, and this was used in the plastic analysis.

The moment versus deflection response and the development of slip with loading as obtained from
the FE analysis are compared with the results from the 2-09 test beam in Figure 7.11. As shown,
the agreement of the FE model with test is good, and the model can be used with confidence for
the purposes of this study.

108
600 600
500 500

Moment (kNm)

Moment (kNm)
400 400
Mpl
300 Mpl 300
FE
200 FE 200
Test-side A
100 Test 100
0 0 Test-side B
0 50 100 0 5 10 15
Deflection (mm) Slip (mm)

(a) Moment versus deflection (b) Moment versus end slip

Figure 7.11 Comparison between FE model and S2-09 test beam response

The effect of the number of shear connectors to be considered when defining the degree of shear
connection η was investigated. For point loaded beams, some interpret the degree of shear
connection as the number of shear connectors over the half span, meaning that when point loads
are present, the shear connectors between the load positions can also be taken into account when
calculating η.

Preliminary FE analyses carried out to validate this suggest that, although the shear connectors on
the intermediate part of the beam (between point loads) seem to be effective, the beneficial effect
they have in reducing slip appears to be relatively small. It was found to range from 15% to 25%
for spans of 10 to 18 m and for point loads at the 1/3 span positions.

In addition to the above observation, the loading pattern in composite beams loaded with point
loads rather than uniform loading plays a beneficial role in terms of shear connection requirements.
FE analyses suggest that shear connection requirements become less severe for point loads at 1/3
span positions in comparison to uniformly loaded beams. The degree of shear connection required
is even smaller for a beam with a single point load at mid-span. These observations are effectively
demonstrated in Figure 7.12 and Figure 7.13.

The case with two point loads at 1/3 span positions is plotted in both graphs and relates to the
number of shear connectors used to estimate η and Mpl. In Figure 7.12, the required number of
studs to satisfy ηmin is placed between the support and the position of the point load. However, in
Figure 7.13 the same number of shear connectors is distributed between the support and mid-
span, which means that the degree of shear connection is approximately 33% lower, but Mpl is also
7 to 8% lower. The slip increased from 0.9 mm to approximately 4 mm in the second case.

1400
1200
Moment [kNm]

1000
800
600
400
200
0
0 1 2 3 4 5 6 7 8
Slip [mm]

Point load @ mid-span Point loads @ 1/3 span positions


UDL 0.95Mpl

Figure 7.12 Comparison of 12 m span composite beam loaded with uniform loading versus point
loads at different positions along the span

109
1400

1200

Moment [kNm] 1000

800

600

400

200

0
0 1 2 3 4 5 6 7 8
Slip [mm]

Point loads @ 1/3 span positions 0.95Mpl

Figure 7.13 12 m span composite beam with a) point loads at 1/3 span positions and b) uniform
loading

7.6 Beams loaded with heavy loads


In some applications. The current minimum degree of shear connection requirements were
developed assuming steel sections than are lighter the those required to support heavier imposed
loads. Composite beams subject to uniform loading have a span/depth ratio of the steel section in
the range of 22 to 27. For more heavily loaded beams, the steel section will also be heavier and
will have a span/depth ratio nearer to 20. FE analyses have shown that the use of a heavier or
deeper steel section necessitates the use of more onerous rules for the minimum degree of
connection (this is analogous to when higher grade steel is used).

Guidance on heavily loaded beams is given in SCI P405. Figure 7.14 compares the degree of
connection for both of loading intensity and 6 mm allowable slip according to Eurocode 4 and NCCI
PN002a-GB. The differences are most noticeable for long span propped beams.

Figure 7.14 Minimum degree of shear connection for 6mm allowable slip based on EN 1994-1-1
and NCCI PN002a-GB

110
7.6.1 Concluding remarks for WP6
This work package has led to the following proposals:
 Current minimum degree of shear connection rules in EN 1994-1-1 are too conservative for
un-propped composite beams that are designed to be not fully utilised in bending. Results
from the current study suggest that a multiplication factor equal to (MEd/MRd)2 applied to the
EN 1994-1-1 shear connection formulae would be suitable (for configurations with proven
ductility of the shear connectors, i.e. slip capacity of at least 6 mm).
 A further modification factor of 0.85 may be applied to the minimum degree of shear
connection for un-propped beams to take account of the loading that is applied to the steel
beam during construction in comparison to propped beams.
 Shear connectors with slip capacities higher than 6mm may lead to lower degrees of shear
connection. A new class of super-ductile shear connectors with a slip capacity of 10mm is
defined.
 The minimum degree of shear connection may be further reduced for cellular beams with
regular circular openings based on the long span beam test in WP3 and finite element
modelling of a wide range of cellular beams
 Additional deflections due to increased slips in composite beams with lower degrees of shear
connection should be taken into account at the serviceability limit state.
 Special considerations will need to be made in terms of the shear connection in composite
beams loaded by heavy off-centre concentrated loads.

7.6.2 Exploitation and impact of the research results


The results of this Work Package form the basis of proposals made to the projects teams involved
with the revisions to Eurocode 4 and so will be of direct relevance to design of composite
construction.

111
112
8 WP7: Project management and coordination

8.1 Objectives of WP7


To manage and coordinate the project and maintain adequate lines of communication between all
the partners involved in the project in order to achieve the project objectives within the time and
budget

To prepare the semester, mid-term and final project reports and financial statements.

8.1.1 Comparison of initially planned activities and work accomplished


All deliverables were prepared although some of them were completed up to 4 months after the
end of the 3 year research. The deliverables were prepared to a common template although
considerable editorial work had to be done to achieve uniformity of style and content.

A total of 8 meetings was held in the course of the research and the following 6 months, including
one combined meeting with ECCS Committee TC 10.

8.1.2 Description and impact of the research results


Composite construction is one of the most important forms of construction for the steel sector as it
leads to materials efficiency and to improved structural performance. Eurocode 4 is recognised as
being one of the most important standard in so far as it provides useable design methods to
encourage use of composite construction. However, there are areas in which Eurocode 4 needs
urgent improvement, and a key area is that of the rules for partial shear connection.

The results of this project DISCCO will be readily exploited as follows:

 By incorporation of improved application rules in Eurocode 4, which is going through its 10 year
revision.
 By improved methods of tests on shear connectors, which lead to more realistic shear connector
behaviour, particular in relation to deformation capacity. This will encourage up-take by
industry.
 By new methods of analysing cellular beams.
 By new empirical factors to take account of the influence of deck shape on shear connector
resistance.
 By improved rules for serviceability performance of composite beams.
SCI and other parties are preparing design guides based on the results of the project which will
show by worked examples how the proposed rules are implemented.

A large number of technical papers have been prepared (6 to 8 is envisaged) on both the testing
work and its analysis, and the proposed design rules.

113
114
9 Exploitation Plan

Composite construction is one of the most important technologies for the steel sector, as it is
widely used in all types of multi-storey buildings particularly those with long spans of 12 to 20 m.
EN1994-1-1 Eurocode 4 has been successful in facilitating the use of composite construction across
Europe but had one major problem in that the shear connection rules were not well adapted to
modern forms of construction.

Therefore, the DISCCO project was aimed at developing more realistic rules for shear connection
behaviour and degree of shear connection. This is timely because Eurocode 4 is going through its
first major review and revision, and so the results of this project will be fed into the project teams
working on these reviews.

The areas in which improvements have been made in DISCCO are:

 New method of shear connector tests involving transverse loading of a percentage of the
applied shear load which leads to improved and more realistic behaviour.
 Shear connector results and modelling for deeper deck profiles (80 mm deep), where different
modes of failure occur, which is a function of the embedment ratio (hsc/hp) of the shear
connector height.
 Development of advanced finite element model using idealised shear connector behaviour.
 Reduction in the minimum degree of unpropped shear connection for composite beams which
cover by far the majority of design cases.
 Models for cellular beams with regular circular openings, and development of shear connection
rules for cellular beams.
 New formula for the shear connector resistance of edge beams without transverse U-bar
reinforcement.
 Development of rules for partially loaded beams, which is generally the case for long span
beams controlled by deflection limits.
The individual exploitation plans are presented under each WP. However, it is important that the
steel sector takes advantage of the results of DISCCO to be able to market the advantages of
modern composite construction. For example, it is now possible to provide practical solutions for
long span asymmetric cellular beams that is not possible in the current Eurocode 4 rules because
sufficient numbers of shear connectors cannot be placed along the beam to satisfy the minimum
shear connection limit.

It may also be possible to develop new deck profiles and shear connector systems that utilise the
results of this project, and which optimises the performance of composite beams.

115
116
References

American Institute of Steel Construction, AISC 360-10, Specification for Structural Steel Buildings,
Manual of Steel Construction, 2010

Aribert, J-M.(1997): ‘Analyse et Formulation Pratique de l’Influence de la Nuance de l’Acier du


Profile sur le Degree Minimum de Connexion Partielle d’une Poutre Mixte’, (Analysis and practical
formula for the grade of steel on the minimum degree of shear connection of composite beams’
Construction Métallique, No. 3, 1997, pp 39-55

BS 5950: Structural Use of Steelwork in Buildings, Part 3 Design in Composite Construction.


Section 3.1 Code of Practice for Design of Simple and Continuous Composite Beams, 1990
including Amendments 2010, British Standards Institution,

CEB-FIP model code 1990. Design code (1993). London: Thomas Telford.

Couchman G (2015), Minimum Degree of Shear Connection Rules for UK Construction to Eurocode
4, Electronic Publication 405, 2015, The Steel Construction Institute, Ascot, UK.

DIN 18800-5: Stahlbauten – Teil 5: Verbundtragwerke aus Stahl und Beton – Bemessung und
Konstruktion.’ Composite construction using steel and concrete -Design and construction’, 2007

EN 1990:2002+A1:2005 (Incorporating corrigendum 2009) Eurocode: Basis of structural design

EN 1992-1-1, Eurocode 2: Design of concrete structures - Part 1-1 : General rules and rules for
buildings + AC:2010, 2004.

EN 1993-1-1: Eurocode 3 Design of Steel Structures Part 1.1 General Rules and Rules for
Buildings, 2005

EN 1994-1-1:2004 (Incorporating corrigendum 2009) Eurocode 4. “Design of composite steel and


concrete structures – Part 1-1: General rules and rules for buildings”, 2005

EN ISO 13918, Welding — Studs and ceramic ferrules for arc stud welding, 2008.
Konrad M.: Tragverhalten von Kopfbolzen in Verbundträgern bei senkrecht spannenden
Trapezprofilblechen ( Structural behaviour of shear connectors in composite beams with
perpendicular orientated trapezoidal sheets) , University of Stuttgart, Institute of Structural Design,
Dissertation, 2011

EN ISO 14555: Welding — Arc stud welding of metallic materials 2014.

Johnson, R.P. & Molenstra (1991), ‘Partial shear connection in composite beams for buildings’,
Proceedings of the Institution of Civil Engineers, Part 2, Vol. 91, December 1991.

Lam D, Elliott K.S., and Nethercot D.A, (2000) Designing composite steel beams with precast
concrete hollow-core slabs, Proc Inst. Civil Engs, Structures and Buildings May 2000, 140, p139-
149

Lungershausen, H (1988): Zur Schubtragfähigkeit von Kopfbolzendübeln. (Shear resistance of


headed shear connectors), Univ., Dissertatation--Bochum, 1988. Bochum: Inst. für Konstruktiven
Ingenieurbau.

Nellinger S, (2015) On the Behaviour of Shear Stud Connections in Composite beams with Deep
Decking’, Disssertation, University of Luxembourg, 2015, PhD-FSTC-2015-53

Nethercot D.A, Li T.Q., Ahmed B, (1998) Plasticity of Composite Beams at the Serviceability Limit
State, The Structural Engineer Vol 76, 12, 1998, p 289-293.

Newmark N.M, Siess C. P and Viesst I.M. (1951) Tests and Analysis of Composite Beams with
Incomplete Shear Connection, Proc. Soc. Experimental Stress Analysis 9 (1), 1951

Qureshi, J., Lam, D. and Ye, J. (2011a), ‘Effect of Shear Connector Spacing and Layout on the
Shear Connector Capacity in Composite Beams’ Journal of Constructional Steel Research, Vol.
67(4), 2011, pp 706-719

Qureshi, J, Lam, D. and Ye, J (2011b). ‘The Influence of Profiled Sheeting Thickness and Shear
Connector's Position on Strength and Ductility of Headed Shear Connector’, Engineering Structures,
Vol. 33(5), 2011, pp 1643 -1656

Ranzi G and Zona A (2007), A Steel-Concrete Composite Beam Model with Partial Interaction
Including Shear Deformability of the Steel Component, Engineering Structures 29, 2007, p 3026 –
3041

117
Ranzi G, Bradford M, and Uy B (2003), A General Method of Analysis of Composite Beams with
Partial Interaction, Steel and Composite Structures, 3, p 169-184

The Steel Construction Institute (2010), NCCI: Modified limitation on partial shear connection in
beams for buildings, PN002a-GB 2010, The Steel Construction Institute, Ascot, UK

Wright HD and Francis RW (1990), Tests on Composite Beams with Low Levels of Shear
connection, The Structural Engineer Vol 68 Issue 15 p 293-298

Background papers used in WP1


Al-deen S, Ranzi G, Vrcelj Z. (2011). Full-scale long-term and ultimate experiments of simply
supported composite beams with steel deck. Journal of Constructional Steel Research, 67, pp.
1658-1676.

Androutsos C, Hosain MU. (1992). Composite beams with headed studs in narrow ribbed metal
deck. In: Easterling WS, Roddis WMK editors, Composite Construction in Steel and Concrete II,
Proceedings of the 2nd International Conference, pp. 771-782.

Bradford M A, Filonov A, Hogan TJ, (2006). Push testing procedure for composite beams with deep
trapezoidal slabs. In: Eleventh international conference on metal structures.

Bradford M A, Filonov A, Hogan TJ, Ranzi G, Uy B, (2006). Strength and ductility of shear
connection in composite T-beams with trapezoidal steel decking. In: Eighth international
conference on steel, space and composite structures, pp. 15-26.

Bradford M A, Pi Y-L, Uy B. 2008, Ductility of composite beams with trapezoidal composite slabs,
In: Composite Construction in Steel and Concrete VI, pp. 151-158.

DIN 2093: Tellerfedern - Qualitätsanforderungen - Maße (Disc springs quality specification), Beuth
Verlag GmbH, 2013-12.

DIN 2092:2006-03: Tellerfedern - Berechnung (Disc springs Calculations), Beuth Verlag GmbH,
2006.

Easterling W S, Gibbings D R, Murray T M (1993). Strength of shear studs in steel deck on


composite beams and joists, Engineering Journal, AISC, 30(2), pp. 44-55.

Ellobody, E; Young, B) (2005). Performance of shear connection in composite beams with profiled
steel sheeting. In: Journal of Constructional Steel Research 62 (7), 2005

Ernst S, Patrick M, Bridge R, Wheeler A. (2005). Reinforcement requirements for secondary


composite beams incorporating trapezoidal decking, In: Leon RT, Lange J editors, Composite
Construction in Steel and Concrete V, Proceedings of the 5th International Conference, pp. 236-
246.

Ernst S, Bridge R, Wheeler A. (2009). Push-out tests and a new approach for the design of
secondary composite beam shear connections. Journal of Constructional Steel Research, 65, pp.
44-53.

Ernst S, Bridge R, Wheeler A. (2010). Correlation of beam tests with push-out tests in steel-
concrete composite beams. ASCE Journal of Structural Engineering, 136(2), pp. 183-192.

Kim B, Wright HD, Cairns R. (2001). The behaviour of through-deck welded shear connectors: an
experimental and numerical study. Journal of Constructional Steel Research, 57, pp. 1359-1380.

Loh H Y, Uy B, Bradford M A. (2004). The effects of partial shear connection in the hogging
moment regions of composite beams: Part I – Experimental study. Journal of Constructional Steel
Research, 60(6), pp. 897-919.

Mottram, J.T. & Johnson, R.P. (1990), ‘Push tests on studs welded through profiled sheeting’, The
Structural Engineer, Vol. 68, No. 10, 1990.

Rackham JW. (2008). Test frame for push specimens. SCI Report RT1227, The Steel Construction
Institute.

Rambo-Roddenberry M. (2002), Behaviour and strength of welded stud shear connectors, PhD
Thesis, Virginia Polytechnic Institute and State University.

Smith A L. (2009a), Resistance of headed stud shear connectors in modern trapezoidal profiles.
SCI Report RT1309.

118
Smith A L. (2009b). Effect of key variables on shear connector performance using new push rig.
SCI Report RT1236.

Smith A L, Couchman GH. (2010). Strength and ductility of headed stud shear connectors in
profiled steel sheeting. Journal of Constructional Steel Research, 66, pp. 748-754.

Books and Publications

Alejano L-R.; Bobet, A, Drucker–Prager Criterion in Rock Mechanics, Rock Eng 45 (6), 2012, p
995–999.

Bode, H (1998): Euro-Verbundbau. Konstruktion, Berechnung. (Euro- Composite construction:


Calculations), 2. Aufl. Düsseldorf: Werner.

Cornelissen H A W, Hordijk D A, Reinhardt H W (1986), Experimental determination of crack


softening characteristics of normal weight and lightweight concrete. Heron 1986;31(2):45–56.

Eggert, F; Nellinger, S; Kuhlmann, U; Odenbreit, C: (2014) Push-out tests with modern deck
sheeting to evaluate shear connector resistances). In: Eurosteel 2014, 7th European Conference on
Steel and Composite Structures, Napoli, Italy (2014) S. 519–520, 10.-12. September 2014.

Jenisch, F. M. (2000): Einflüsse des profilierten Betongurtes und der Querbiegung auf das
Tragverhalten von Verbundträgern, (Influences of the profiled concrete section and transverse
bending on the structural behaviour of composite beams), Fachbereich Architektur, Raum- und
Umweltplanung, Bauingenieurwesen, Universität Kaiserslautern, Dissertation, 2000

Kuhnle, T (2012) Untersuchung zur Annahme der Materialkennwerte bei der Nachrechnung von
Verbundversuchen mit der Finite-Element-Methode (Research into the properties of materials in the
behaviour of composite construction using the finite lement method), Institut für Konstruktion und
Entwurf im Stahl-Holz-und Verbundbau, Universität Stuttgart, 2012

Laulusa, A.; Bauchau, O. A.; Choi, J-Y.; Tan, V.B.C.; Li, L. (2006): Evaluation of some shear
deformable shell elements. In: International Journal of Solids and Structures 43 (17), S. 5033–
5054. DOI: 10.1016/j.ijsolstr.2005.08.006.

Martin Gertis (2013): Numerische Analyse der Tragfähigkeit von CoSFB-Verbundträgern.


(Numerical analysis of the resistance of CoSFB composite beams) Nr. 2013-39X Gertis.
Diplomarbeit. Universität Stuttgart, Stuttgart. Institut für Konstruktion und Entwurf.

Roik, K.; Lungershausen, H (1989).: Zur Tragfähigkeit von Kopfbolzendübeln in Verbundträgern


mit unterbrochener Verbundfuge (Trapezprofildecken). Load bearing capacity of studs in composite
beams with an interrupted connection ( trapezoidal decking) , In: Stahlbau 58 9 (1989), S. 269-
273

Roik, K (1993): Verbundkonstruktionen. Bemessung auf der Grundlage des Eurocode 4, Teil 1-1
(Composite construction assesment on the basis to Eurocode 4-1-1)); Sonderdruck aus dem
Betonkalender (Special edition of Betonkalender) 1993. Berlin: Ernst.

Weischedel, B: Numerische Untersuchungen an Push-out Prüfkörpern zum Tragverhalten von


profilierten Verbundquerschnitten (Numerical research on the push specimens into profiles section
properties) , Institut für Konstruktion und Entwurf, Masterarbeit, Universität Stuttgart, 2015

Werkle, H (2008): Finite Elemente in der Baustatik. Statik und Dynamik der Stab- und
Flächentragwerke. 3., aktualisierte und erweiterte Auflage (Finite elements in structures)
Wiesbaden: Friedr. Vieweg & Sohn Verlag GWV Fachverlage GmbH, Wiesbaden.

3DS - Simulia (2012): Getting Started with Abaqus. Interactive Edition. Abaqus 6.12. Providence,
RI, USA.

119
120
10 LIST OF FIGURES

Figure 2.1 Shear connector resistances from tests compared to characteristic resistance to
EN 1994-1-1 9
Figure 2.2 Shear connector resistances obtained from tests by SCI compared with characteristic
resistance to EN 1994-1-1 (C=central position, F=favourable position, TS=top of slab,
TD= top of deck) 10
Figure 2.3 Slips obtained from tests by the SCI compared with EN 1994-1-1 requirement.
Transverse loading equivalent to 12% of the test load 10
Figure 2.4 Forces acting on shear connectors in a composite beam 11
Figure 2.5 Test set-up for the application of transverse loading 12
Figure 2.6 Details of the trapezoidal deck profiles 12
Figure 2.7 Details of the trapezoidal deck profiles 13
Figure 2.8 Test setup used at the University of Stuttgart 13
Figure 2.9 Typical dimensions of specimens with Cofraplus 60 decking 13
Figure 2.10 Typical dimensions of specimens with ComFlor 80 decking 14
Figure 2.11 Typical dimensions of specimens with Cofrastra 56 decking 14
Figure 2.12 Dimensions of specimens with steel decking parallel to the beam 15
Figure 2.13 Typical load-slip curve for solid slab (specimen 1-01-1) 15
Figure 2.14 Typical load-slip curve for specimen with CP 60 decking (specimen 1-04-1) 19
Figure 2.15 Failure surface for concrete pull-out for CP60 decking 19
Figure 2.16 Stud deformation with 2 plastic hinges for CP60 decking 19
Figure 2.17 Typical load-slip curve for specimen with CF80 decking (specimen 1-10-3) 20
Figure 2.18 Failure surface for rib pry-out failure for CF 80 decking 20
Figure 2.19 Stud deformation with one plastic hinge for CF 80 decking 20
Figure 2.20 Typical load-slip curve for specimen with CF56 decking (specimen 3-03-1) 21
Figure 2.21 Load-slip curves for specimens with parallel deck ribs in series KL 21
Figure 2.22 Load-slip curves with and without tension ties for specimens using CF80 decking 23
Figure 2.23 Influence of concentric transverse loading using CP 60 decking 25
Figure 2.24 Influence of concentric transverse loading for CF 80 with pairs of studs per rib 25
Figure 2.25 Influence of concentric transverse loading using CF80 and single studs per rib 25
Figure 2.26 Influence of eccentric transverse loading using CP 60 decking 26
Figure 2.27 Influence of eccentric loading using CF80 decking 26
Figure 2.28 Influence of the deck shape (comparison of tests 2-01-1 and 3-03-1) 27
Figure 2.29 Influence of number of studs per rib for rib pry-out failure (comparison of tests 1-10-3
(pairs of studs) and NR1-1 (single studs)) 27
Figure 2.30 Influence of the diameter of the shear connectors 28
Figure 2.31 Influence of stud welding procedure using CP 60 (tests 2-01-1 and 2-03-1) 28
Figure 2.32 Influence of the stud welding procedure using CF 80 (tests 1-10-3 (through-deck
welded) and 3-02 (pre-punched deck)) 29
Figure 2.33 Influence of reinforcement pattern (comparison of one and two layers) 29
Figure 2.34 Influence of reinforcement pattern on rib failure (tests 1-10-3 and 3-01-3)) 30
Figure 2.35 Influence of concrete strength on shear connector resistance 30
Figure 2.36 Comparison of "characteristic" test results to predicted characteristic resistance for
characteristic material properties 32
Figure 2.37 Minimum width of concrete slab for push-out tests 32
Figure 2.38 Dimensional and detailing requirements for push-out test 33
Figure 2.39 Geometric parameters used in the new model 34
Figure 2.40 Statistical analysis of the model of Nellinger for 375 push- tests 35
Figure 2.41 Statistical analysis of the model of Nellinger in its range of application 35
Figure 3.1 Test setup for beams with CP60 decking, series 2-01 – 2-08 (U Stuttgart) 38
Figure 3.2 Test setup for beams with CF80 decking series 2-09 – 2-10 (U Luxembourg) 38
Figure 3.3 Test setup for CF80 decking series 2-09 – 2-10 (U Luxembourg) 39
Figure 3.4 Test setup for CF80 decking series 2-09 – 2-10 (U Luxembourg) 39
Figure 3.5 Load-deflection curve for beam 2-01 40
Figure 3.6 Test setup of beam 2-02 40
Figure 3.7 Test 2-02 at failure 40
Figure 3.8 Load-deflection curve for beam 2-02 40
Figure 3.9 Load-deflection curve for beam 2-03 41
Figure 3.10 Load-deflection curve for beam 2-04 41
Figure 3.11 Load-deflection curve for beam 2-05 42
Figure 3.12 Load-deflection curve for beam 2-06 42
Figure 3.13 Load-deflection curve for beam 2-07 43
Figure 3.14 Load-deflection curve for beam 2-08 43
Figure 3.15 Test setup of specimen 2-09 at the beginning of the test 44
Figure 3.16 Test 2-07 at maximum mid-deflection 44
Figure 3.17 Load-deflection curve for beam 2-09 44
Figure 3.18 Load-deflection curve for beam 2-10 44
Figure 3.19 Measurement of slip along the beam 45
121
Figure 3.20 Slip development for beam test 2-01 45
Figure 3.21 Results cutting specimen 2-01 on both ends of the beam 45
Figure 3.22 Strain in the lower flange of the steel beam for beam test 2.01 46
Figure 3.23 Results of the plastic resistance of the beam tests using the characteristic shear
connector resistance (Pk= 0.9 Ptest or Pk to EC4) 48
Figure 4.1 Cross-section dimensions of the asymmetric cellular beam 50
Figure 4.2 Set-up for uniformly distributed load test on cellular beam 50
Figure 4.3 Set-up for uniformly distributed load test on cellular beam 51
Figure 4.4 Applied load versus mid-span deflection for cellular beam test 51
Figure 4.5 Applied load versus end-slip measured by LVDT_11 for cellular beam test 52
Figure 4.6 Applied load versus strain in bottom flange at mid-span for cellular beam test 52
Figure 4.7 Testing arrangement for asymmetric load test 53
Figure 4.8 Applied load versus deflection at loading position for the asymmetric load test 53
Figure 4.9 Deflected shape of the beam and elongated opening for the asymmetric load test 53
Figure 4.10 Load versus deflection difference across the opening from LVDT_4 to LVDT_6 in the
shear test 54
Figure 4.11 Strain distribution through cross-section depth corresponding to measured strains for
cellular beam test 1 54
Figure 4.12 Arrangement of actuators and load-spreader beams for the 11.2m span fabricated
beam test 55
Figure 4.13 Deflection of composite fabricated beam after unloading from failure 56
Figure 4.14 Relationship between load and mid-span deflection for fabricated beam 56
Figure 4.15 Relationship between load and end-slip for fabricated beam 56
Figure 4.16 Relationship between load and strain in bottom flange for solid web beam test 57
Figure 4.17 Strain distribution through cross-section corresponding to measured strains and
average measured flange strains recorded for the asymmetric fabricated beam 57
Figure 4.18 Load-deflection cycle for the asymmetric fabricated beam up to twice working load and
comparison with the theory 58
Figure 5.1 Beam arrangement and floor plate dimensions 62
Figure 5.2 Layout of beams in floor plate test 63
Figure 5.3 View of load test on secondary beam 64
Figure 5.4 Details for shear connectors at primary beam and edge beam 64
Figure 5.5 Details of U-bar arrangement at edge beam 65
Figure 5.6 Deflection (linear voltage displacement transducers) positions LVDT 1 to 16 65
Figure 5.7 Deflection of secondary beams close to failure 66
Figure 5.8 Load-displacement graph for internal secondary beams up to failure 67
Figure 5.9 Load-displacement graph for internal primary beam up to failure 67
Figure 5.10 View of load test on edge beam 68
Figure 5.11 Load -displacement graph for the edge beam test without U bars 69
Figure 5.12 Edge beam test without U bars at an equivalent loading of 37 kN/m2 69
Figure 5.13 Deflected shape of edge beam test with U bars close to failure 69
Figure 5.14 Load -displacement graph for the edge beam test with U bars 70
Figure 5.15 Deflection of edge beam due to missing support to central column 70
Figure 5.16 Deflection of column with a missing support subject to an equivalent uniform loading
acting on the edge beams 71
Figure 5.17 Deformation of floor plate with loads on the internal secondary beams 72
Figure 5.18 Comparison of FE modelling against test results for internal beam test 72
Figure 5.19 Deformation of composite plate with loads on the edge beams 73
Figure 5.20 Comparison of FE model prediction the tests on edge beams 73
Figure 6.1 Load bearing behaviour of a headed stud in a solid slab, Lungershausen 75
Figure 6.2 Behaviour of headed studs in profiled slabs according to Lungershausen 76
Figure 6.3 ABAQUS model for shear connectors in composite slabs 77
Figure 6.4 Concrete behaviour in compression 78
Figure 6.5 Stress-crack width and tension damage relationship for concrete 78
Figure 6.6 Comparison of test results and ABAQUS results for test1-03-2 79
Figure 6.7 Deformation of the headed studs within the concrete slab 79
Figure 6.8 Plastic strains of the concrete slab obtained from the FE model 80
Figure 6.9 Concrete damage and concrete pull-out failure of test 1-03-3 compared to ABAQUS
model 80
Figure 6.10 Steel and concrete properties used in the FE models 82
Figure 6.11 Comparison of FE model and 15m span cellular beam test in WP3 83
Figure 6.12 Prediction of the performance of the 11m span asymmetric beam test using ABAQUS83
Figure 6.13 Description of the longitudinal shear connection model in ABAQUS 84
Figure 6.14 Distribution of Von Mises stresses in the beam at failure 85
Figure 6.15 Comparison between mid-span deflections for the 15.3m span cellular beam test and
obtained using ABAQUS 85
Figure 6.16 Shear connector properties assumed in the model of the cellular beam 86
Figure 6.17 Deformed shape (left) and strain in mid-span at the elongated cell (right) from the
analysis of the 15.3 m span cellular beam test using ANSYS 86
122
Figure 6.18 Comparison between FE modelling and test results for 15.3 m span cellular beam using
ANSYS 86
Figure 6.19 Comparsion of strains in FE model and test results for 15.3 m span cellular beam 87
Figure 6.20 Minimum degrees of shear connection for unpropped cellular beams 88
Figure 6.21 Comparison between current rules, proposed modification to the rules and ANSYS FE
results for asymmetric cellular beams of 9 m, 12 m and 15 m span 90
Figure 6.22 Simplified model using 1D and 2D elements for the slab and steel section 92
Figure 6.23 Comparison of simplified FEA model versus test beam 2-10 of WP2 92
Figure 6.24 Beam sizes used in the study and idealised load-slip behaviour of the shear
connectors 92
Figure 6.25 Results of simplified FE analysis of 6 to 12m span composite beams 93
Figure 7.1 Minimum shear connection rules in EN 1994-1-1 for symmetric and asymmetric beams
and S275 and S355 steel grades 95
Figure 7.2 Forces and displacements in a composite beam as affected by slip 96
Figure 7.3 Load versus deflection for long span beams up to twice working load and comparison
with the elastic theory 98
Figure 7.4 Load versus deflection curves obtained from FEA and tests on the 11m span
asymmetric beam in WP3 103
Figure 7.5 Parametric study results for single shear connector per deck rib and 6 mm slip capacity
(UR = Utilisation ratio = M/Mpl) 104
Figure 7.6 Parametric study results showing the effect of different deck/slab and shear connector
dimensions (12 m span case) 105
Figure 7.7 Parametric study results showing the effect of beam asymmetry (Afb/Aft=3) 105
Figure 7.8 Results from the University of Luxembourg in terms of utilisation factor and slip and
degree of shear connection 106
Figure 7.9 Graph showing the effect of utilisation factor on the minimum degree of shear
connection for 6mm slip 106
Figure 7.10 Test setup for series 2-09 and 2-10 – side view 108
Figure 7.11 Comparison between FE model and S2-09 test beam response 109
Figure 7.12 Comparison of 12 m span composite beam loaded with uniform loading versus point
loads at different positions along the span 109
Figure 7.13 12 m span composite beam with a) point loads at 1/3 span positions and b) uniform
loading 110
Figure 7.14 Minimum degree of shear connection for 6mm allowable slip based on EN 1994-1-1
and NCCI PN002a-GB 110

123
124
11 LIST OF TABLES

Table 2.1Data for push-out tests in Task 1.2 16


Table 2.2Data for push-out tests in Task 1.3 17
Table 2.3Parameters of tests with the decking placed parallel to the beam 18
Table 2.4Push-out test results for CF 80 decking in Task 1.2 22
Table 2.5Push-out test results for CF 80 decking in Task 1.2 24
Table 3.1Beam tests and their parameters performed in Task 1.3 38
Table 3.2Tensile test results (yield point) for IPE sections used in WP 2 39
Table 3.3Results of concrete compression and tensile tests 39
Table 3.4Shear connector resistances Pstud [kN] as input in the plastic analysis of the beam
tests (ST) 47
Table 3.5 Model factor (MTest/MR,calc) for the beam tests in WP2 47
Table 4.1 Comparison of the test results on long span beams with Code methods 59
Table 5.1 Summary for test on primary beam with loaded internal secondary beams allowing for
10% load spread to each of the edge beams 67
Table 5.2 Summary for test on edge beams with and without U bars 69
Table 6.1 Material parameters concrete damaged plasticity 77
Table 6.2 Change of the stud shear resistance with various parameters 81
Table 6.3 Change of the stud shear resistance with various parameters (initial value D22x125) 81
Table 6.4 Change of the stud shear resistance (initial value C30/37) 81
Table 6.5 Summary of geometrical data for configurations in the parametric analysis 87
Table 6.6 Minimum degree of shear connection for cellular beams in comparison to the
equivalent solid web beam based on FEA 89
Table 6.7 Asymmetric cellular composite beam cases considered 90
Table 6.8 Comparison of deflection of composite solid web beams and cellular beams and
comparison with approximate equation 91
Table 7.1 Summary of shear connector stiffness from push tests (values rounded) 97
Table 7.2 Data of the composite beams used in the FE models 99
Table 7.3 Deflection of composite beams at 30% and 50% of their maximum load, Pu
corresponding to the bending resistance Mu based on finite element results 99
Table 7.4 Comparison of deflections and the elastic theory in the FE models 99
Table 7.5 Comparison of deflections of solid web beams and cellular beams from FEA, theory for
rigid and flexible shear connectors 100
Table 7.6 Minimum degree of shear connection for symmetric beams based on 1mm slip limit at
the serviceability limit state 101
Table 7.7 Minimum degree of shear connection for asymmetric beams based on original IPE
section with flange areas of 3:1 based on 1 mm slip limit at the serviceability limit
state 101
Table 12.1 Acronyms and abbeviations 127
Table 12.2 Nomenclature 127

125
126
12 LIST OF ACRONYMS AND ABBREVIATIONS

Table 12.1 Acronyms and abbeviations

Acronym Abbreviation
AM ArcelorMittal

EC4 Eurocode 4 or EN 1994-1-1


FEA Finite Element Analysis
LVDT Linear Voltage Displacement Transducer
SCI Steel Construction Institute

UBrad University of Bradford

ULux University of Luxembourg


UStutt University of Stuttgart

UR Utilisation Ratio in bending of composite beam at the ultimate limit


state
WP Work Package

Table 12.2 Nomenclature

Symbol Unit Meaning


Afb [mm2] Area of bottom flange of beam
Aft [mm2] Area of top flange of beam
dsc [mm] Stud diameter
η [-] Degree of shear connection
ηmin [-] Minimum degree of shear connection
fc,dry,cyl [N/mm2] Concrete strength measured by cylinders stored in water until an
age of 7 days
fy [N/mm2] Yield strength of steel used in the beam
fu [N/mm2] Tensile strength of shear connector
h [mm] Height of steel beam
ho [mm] Height of opening in beam web
hp [mm] Height of deck profile
hs [mm] Height of concrete slab in push-out test
hsc [mm] Height of studs after welding
k kN/mm Shear stiffness of shear connectors
kt [-] Reduction factor for the resistance of shear connectors in
transversely spanning composite deck ribs
Le [mm] Effective span of beam between points of zero moment
MEd [kNm] Applied moment on composite beam
MPl,a [kNm] Bending resistance of steel section
MPl,Rd [kNm] Bending resistance of composite beam for full shear connection
MRd [kNm] Bending resistance of composite beam for partial shear connection
no [-] Number of openings in beam span
nr [-] Number of studs in a row
ped [kN/m] Distributed transverse load applied to push-out test
P0.9min [kN] 90% of the minimum static load per stud in a push-out test series
Pe [kN] Load per stud in push-out test
Pe,static [kN] Static load per stud in push-out test

127
PR [kN] Stud resistance obtained from test
PRk [kN] Characteristic stud resistance obtained from test = 0.9 PR for single
test
PRm [kN] Mean value for stud resistance
PRd [kN] Design value for stud shear resistance
Pr,Ed [kN] Transverse load per row of shear connectors
PT [kN] Transverse load applied by hydraulic press to push-out test
PV [kN] Vertical load applied to push-out test
s [mm] Slip in the shear connectors
ssc [mm] Longitudinal spacing of shear connectors
t [mm] Thickness of steel in composite deck
wadd [mm] Additional displacement of beam due to openings
wb [mm] Displacement of beam without openings

128
HOW TO OBTAIN EU PUBLICATIONS

Free publications:
• one copy:
via EU Bookshop (http://bookshop.europa.eu);
• more than one copy or posters/maps:
from the European Union’s representations (http://ec.europa.eu/represent_en.htm);
from the delegations in non-EU countries (http://eeas.europa.eu/delegations/index_en.htm);
by contacting the Europe Direct service (http://europa.eu/europedirect/index_en.htm) or
calling 00 800 6 7 8 9 10 11 (freephone number from anywhere in the EU) (*).
(*) The information given is free, as are most calls (though some operators, phone boxes or hotels may charge you).

Priced publications:
• via EU Bookshop (http://bookshop.europa.eu).
KI-NA-28-458-EN-N
This Final Report describes the technical activities carried out under this RFCS
contract RFSR CT 2012—00030 which is concerned with developing new rules
for the partial shear connection of composite beams. Based on a comprehensive
series of push tests on deck profiles of 56 to 80 mm height with 19 mm and
22 mm diameter shear connectors, a proposed standard for the push-out test
regime has been prepared which permits use of a transverse loading of 5% of
the applied shear load.

Ten short span beam tests were performed in order to verify that plastic analysis
principles may be used for beams with low degrees of shear connection using
shear connector resistances obtained from the push tests.

A 15.3 m span composite cellular beam with 36% degree of shear connection
was tested in order to examine the extreme case of partial shear connection
and it was found that end slips up to 20 mm could be developed at the plastic
bending resistance of the composite beam for the particular degree of shear
connection. An 11.2 m span composite asymmetric beam with 33% degree of
shear connection was also tested, and again the failure load exactly matched the
predicted bending capacity to Eurocode 4.

As the final part of the testing programme, a 10 m x 4 m floor plate consisting


of 9 beams and 6 columns was tested to investigate the interaction effects
between the secondary and primary beams and also the effect of the
reinforcement around the shear connectors at edge beams.

Numerical studies with regard to modelling of composite beams were used for
predictions related to the short span and long span beam tests. The numerical
analyses also extended to the local behaviour of the shear connectors
particularly when affected by concrete cracking and high strains due to concrete
crushing.

Studies on the relaxation in shear connection rules for composite beams that are
partially utilised in bending and for cellular beams have been performed which
have led to easy to use results.

ISBN 978-92-79-65673-6
doi:10.2777/923858

You might also like