Sunteți pe pagina 1din 38

School of Civil Engineering Sydney NSW 2006 AUSTRALIA http://www.civil.usyd.edu.

au/
Centre for Advanced Structural Engineering

Structural Modelling of Support Scaffold Systems


Research Report No R896

Tayakorn Chandrangsu BSc MSc Kim JR Rasmussen MScEng PhD

June 2009
ISSN 1833-2781

School of Civil Engineering Centre for Advanced Structural Engineering http://www.civil.usyd.edu.au/

Structural Modelling of Support Scaffold Systems


Research Report No R896 Tayakorn Chandrangsu, BSc, MSc Kim JR Rasmussen, MScEng, PhD

June 2009

Abstract:
In this report, accurate three-dimensional advanced analysis models are developed to capture the behaviour of support scaffold systems, as observed in full-scale subassembly tests consisting of three-by-three bay scaffold systems with combinations of various lift heights, number of lifts and jack extensions. The paper proposes methods for modelling spigot joints, semi-rigid upright-to-beam connections and base plate eccentricities. Material nonlinearity is taken into account in the models based on the Ramberg-Osgood expression fitted to available experimental data. Actual initial geometric imperfections including member out-ofstraightness and storey out-of-plumb are also incorporated in the models. The ultimate loads from the nonlinear analyses were calibrated against failure loads and load-deflection responses obtained from full-scale subassembly tests. The numerical results show very good agreement with tests, indicating that it is possible to accurately predict the behaviour and strength of highly complex support scaffold systems using material and geometric nonlinear analysis. The report is a milestone in the ongoing development of a design methodology for support scaffold systems based on advanced analysis currently undertaken at the University of Sydney.

Keywords:
Advanced analysis, Formwork subassemblies, Support scaffold systems, Steel scaffolds, Falsework, Subassembly tests, Structural models, Calibrations

Structural Modelling of Support Scaffold Systems

June 2009

Copyright Notice School of Civil Engineering, Research Report R896 Structural Modelling of Support Scaffold Systems 2009 Tayakorn Chandrangsu and Kim JR Rasmussen T.Chandrangsu@usyd.edu.au and K.Rasmussen@usyd.edu.au ISSN 1833-2781 This publication may be redistributed freely in its entirety and in its original form without the consent of the copyright owner. Use of material contained in this publication in any other published works must be appropriately referenced, and, if necessary, permission sought from the author.

Published by: School of Civil Engineering The University of Sydney Sydney NSW 2006 AUSTRALIA June 2009 This report and other Research Reports published by the School of Civil Engineering are available on the Internet: http://www.civil.usyd.edu.au

School of Civil Engineering Research Report No R896

Structural Modelling of Support Scaffold Systems

June 2009

Table of Contents 1. Introduction.......................................................................................................5 1.1 Scaffold Systems.........................................................................................5 1.2 Advanced Analysis .....................................................................................6 1.3 Previous Scaffold Models ...........................................................................7 2. Full-Scale Subassembly Tests.........................................................................10 2.1 Test Setup and Procedures ........................................................................10 2.2 Test Configurations...................................................................................12 2.3 Test Results ...............................................................................................15 3. Finite Element Models....................................................................................17 3.1 Spigot Joints..............................................................................................17 3.2 Semi-Rigid Standard-to-Ledger Connections...........................................18 3.3 Brace Connections ....................................................................................19 3.4 Base Plate Eccentricity .............................................................................20 3.5 Load Eccentricity ......................................................................................20 3.6 Geometric Imperfections ..........................................................................20 3.7 Geometric and Material Nonlinearities.....................................................21 3.8 Calibrations ...............................................................................................25 4. Discussion .......................................................................................................35 5. Conclusions.....................................................................................................36 Acknowledgement ..............................................................................................36 References ...........................................................................................................36

School of Civil Engineering Research Report No R896

Structural Modelling of Support Scaffold Systems

June 2009

School of Civil Engineering Research Report No R896

Structural Modelling of Support Scaffold Systems

June 2009

1. Introduction
1.1 Scaffold Systems Scaffolds are temporary structures commonly used in construction to support various types of loads. The vertical loads on scaffold can be from labourers, construction equipment, formworks, and construction materials. Commonly, scaffolds must also be designed to withstand lateral loads, including wind loads, impact loads, and earthquake loads. Depending on their use, scaffolds may be categorised as access scaffolds or support scaffolds. Access scaffolds are used to support light to moderate loads from labourers, small construction material and equipment for safe working space. They are usually attached to buildings with ties and only one bay wide. Support scaffolds, also sometimes called falsework, are subjected to heavy loads, for example, concrete weight in the formwork. Support scaffolds are the main focus in this research. An example of a support scaffold system is shown in Figure 1. Support scaffolds normally consist of standards (vertical members), ledgers (horizontal members), and braces. The scaffold standards are connected to each other to create a lift via couplers, also known as spigot joints (Figure 2). In order to connect ledgers to standards, wedge-type or Cuplok joints (Figure 3) are usually preferred for the connection because no bolting or welding is required; though, in some systems manually adjusted pin-jointed couplers are still being used. The connections for diagonal brace members are usually made of hooks for easy assembling. The base of scaffolds consists of threaded adjustable jacks, which can be extended up to typically 600 mm by a wing nut to accommodate irregularity of the ground. The top of scaffolds consists of threaded adjustable jacks with U-heads which support timber bearers and ensure the levelling of the formwork.

Figure 1: Typical support scaffold

School of Civil Engineering Research Report No R896

Structural Modelling of Support Scaffold Systems

June 2009

Top Standard

Spigot Insert

Bottom Standard

Figure 2: Schematic of spigot joint


3 Locking pin 2

Top cup Ledger blade 1

Bottom cup Standard

Ledger

Figure 3: Schematic of Cuplok joint 1.2 Advanced Analysis With the ready availability of powerful computers and sophisticated structural analysis software packages, geometric and material nonlinear structural analysis has become feasible and practical. Nonlinear analysis allows researchers and practitioners to more accurately predict the failure load and deformation of scaffold systems. Advanced analysis involves the modelling of changes of the geometry of structures as a result of loading and inelastic material behaviour. In the research by Gylltoft and Mroz [1], a three-dimensional geometric and material nonlinear finite element model was verified against the results of a full scale test scaffold. The model was further applied to determine the ultimate load of a typical access scaffold considering various configurations and load combinations. However, in many cases research on scaffold systems has focused on elastic nonlinear geometric modelling associated with second-order effects. For example, elastic geometric nonlinear analyses were reported by Peng et al. [2], Prabhakaran et al. [3], Yu et al. [4], Chu et al. [5], and Weesner and Jones [6]. Geometric nonlinear analysis is also a common practice in design offices, whereas the use of inelastic analysis is still rare. Nevertheless, with accurate finite element model, advanced analysis method can usually fulfil the design requirement with no tedious separate member capacity checks.
School of Civil Engineering Research Report No R896

Structural Modelling of Support Scaffold Systems

June 2009

AS4100 [7] allows the application of advanced analysis for the design of steel frames in which the members are of compact cross-section with full lateral restraint, thus preventing local buckling and flexural-torsional buckling. Advanced analysis models should include related material properties, residual stresses, instability effects, initial geometric imperfections, actual connection behaviour, construction methods, and interaction with the foundations [8]. In the research described herein, three-dimensional advanced structural analysis models are proposed in order to develop a new design methodology for support scaffold systems. 1.3 Previous Scaffold Models By means of available commercial finite element softwares such as ANSYS [9], LUSAS [10], and NIDA [11], many new studies on scaffold behaviour were carried out through threedimensional models such as those presented by Prabhakaran et al. [3], Milojkovic et al. [12], and Godley and Beale [13]. Three-dimensional structural analysis is beneficial in describing complex failure modes such as those observed when the combined effects of in-plane and out-of-plane bending are present. Some past models proposed by Huang et al. [14], and Peng et al. [15] were created in two dimensions for simplicity and computational efficiency. Two types of geometrical imperfections are typically required to be considered in an advanced analysis for steel framed systems to capture the second-order effects: the initial member out-of-straightness of the standard and the initial story out-of-plumb of the frame. There are many ways of taking geometric imperfection effects into account. Three methods of modelling imperfections were trialled in [16], including the scaling of eigenbuckling modes (EBM), the application of notional horizontal forces (NHF), and the direct modelling of initial geometric imperfections (IGI). EBM was performed by carrying out an eigenbuckling analysis on the structural model, and then scaling and superimposing the lowest eigenmode onto the perfect geometry to create an initial imperfect structural frame for the second-order structural analysis. In the NHF approach, additional lateral point loads were applied at the top of each column in one direction of the frame and initial member out-ofstraightness could be represented by lateral distributed forces along each member. The IGI method consisted of applying an initial sway of the frame and an out-of-straightness to each column in the frame. For scaffold systems, these same approaches can be applied to model the effects of initial imperfections in the analysis. For example, Yu et al. [4] and Chu et al. [5] integrated EBM with the magnitude of the column out-of-straightness of 0.001 of the height of the scaffold units into the model. Moreover, Yu and Chung [17] investigated a method called critical load approach where initial imperfections were integrated directly into a Perry-Robertson interaction formula to determine the failure loads of the scaffolds in the analysis. In other research on scaffold systems by Peng et al. [2], the NHF approach was incorporated in the model by applying a horizontal notional force of 0.1% to 0.5% of the vertical loads at midheight of the scaffold lift. Godley and Beale [18] adopted an IGI approach by imposing a sinusoidal bow to the members and angular out-of-plumb to the frame. In each of these approaches, careful calibration against test results or numerical reference values is required. Scaffold joints are complex in nature due to need for rapid assembly and reassembly in construction. The Cuplok connections behave as semi-rigid joints, and show looseness with small rotational stiffness at the beginning of loading. Once the joints lock into place under applied load, the joints become stiffer [13]. Wedge-type joints are generally more flexible and closer to pinned connections. They also often display substantial looseness at small rotations [18]. Figure 4 shows typical moment-rotation curves for cuplock [13] and wedgetype [18] joints. As to spigot joints, the spigot can create out-of-straightness of the standards,
School of Civil Engineering Research Report No R896

Structural Modelling of Support Scaffold Systems

June 2009

and the possibility of the joint to open up due to the gap between the standard and the spigot can produce complexity in modelling [19]. Moreover, the modelling of boundary conditions of scaffold systems is crucial because the top and bottom restraints can significantly influence the stability and strength of the system [20]. In recent research by Peng et al. [2], analysis models of wedge-type jointed, 3-storey, 3-bay, and 5-row scaffold system were presented. Experimental tests on scaffold joints showed that the joint stiffness varied between 4.903 kNm/rad (50 ton cm/rad) and 8.826 kNm/rad (90 ton cm/rad) with the average of 6.865 kNm/rad (70 ton cm/rad) being adopted for all joints into their model. Godley and Beale [18] found that scaffold connections are frequently made of wedge-type joints, for which the joint stiffness exhibits different response under clockwise and counterclockwise rotations, and occasionally exhibits looseness in connections with low stiffness. Consequently, Prabhakaran et al. [3] modified the stiffness matrix for the end points of the beam to include connection flexibility, using a piecewise linear curve to model the momentrotation response.

3.5 3.0 Cuplok joint Moment (kNm) 2.5 2.0 1.5 1.0 0.5 0.0 0.00 Wedge-type joint

0.05

0.10 Rotation (radian)

0.15

0.20

Figure 4: Typical moment-rotation curves for Cuplok and wedge-type joints Yu [20] studied the effects of boundary conditions of scaffold systems, and categorised them into four cases, i.e. Pinned-Fixed, Pinned-Pinned, Free-Fixed, and Free-Pinned, with the first term being the translational restraint at the top of the scaffold, and the second term being the rotational restraint at the base of the scaffold. In all analyses, the rotation at the top was assumed to be free. These conditions were incorporated into the models of one bay of onestorey modular steel scaffolds (MSS1), and two-storey modular steel scaffolds (MSS2). Yu found that for MSS1 the failure loads for Free-Fixed and Pinned-Pinned conditions are reasonably close to test results; however, for MSS2 the model results are considerably higher than the test results. Subsequently, Yu suggested that since the top of the scaffolds normally has lateral restraints then joints at the top can be modelled as translational springs, and for the bottom rotational spring can be applied. A stiffness of 100 kN/m for the top translational spring and stiffness of 100 kNm/rad for the bottom rotational spring gave comparable results to the tests.

School of Civil Engineering Research Report No R896

Structural Modelling of Support Scaffold Systems

June 2009

Chu et al. [5] studied single storey double bay scaffolds. In the presence of restraints on the loading beam and the jack bases, the top and base were modelled with various boundary conditions, and the scaffold connections were assumed to be rigid. The researchers found that both Pinned-Pinned and Pinned-Fixed conditions gave higher load carrying capacities than the experimental results; on the other hand, the Free-Fixed condition gave satisfactory result compared to the tests. Research on the stability of single storey scaffold systems by Vaux et al. [21] found that when Cuplok connections are represented by pin joints, and the connections of the top and bottom jacks to the standards are assumed as rigid with the topbottom boundary conditions taken as Pinned-Pinned, good agreement can be achieved between numerical and experimental failure loads. Weesner and Jones [6] studied the load carrying capacity of three-storey scaffolds assuming rigid joints between the stories, and pin joints for the top and the bottom boundary conditions. The results of their elastic buckling analysis were higher than the test values with the percentage differences ranging from 6% to 17%. In the analysis of large access scaffold systems by Godley and Beale [18], cantilever arm tests were carried out on scaffold wedgetype joints. The nonlinear moment-rotation curve from the tests showed joint looseness and different values of rotational stiffness under positive (counter-clockwise) rotation and negative (clockwise) rotation. The authors suggested the use of a multi-linear or nonlinear moment-rotation curve for scaffold joint modelling. In the work by Enright et al. [19], the modelling of spigot joints was studied for the stability analysis of scaffold systems. The spigot insert (Figure 2) was considered to have bending resistance, but not to transmit axial load; therefore, the model adopted two vertical members connected by pin joints representing the standards, and on the side, the entirely rigid spigot member was connected at the top, centre, and bottom to the standard via short and axially stiff members capable of transferring only lateral forces, as shown in Figure 5. Due to the axial load in the standards, the spigot would be in bending, and the amount of bending would depend on the amount of axial load and the degree of out of straightness. From research of Harung et al. [22], it was found that if the spigot joints are modelled as fully continuous joints, the analysis would overestimate the load carrying capacity of the system.
Axial load Pin joint Standards Spigot

Figure 5: Spigot joint model Milojkovic et al. [23] studied eccentricity in the modelling of scaffold connections. Given that the neutral axes of the connections were offset by 50 mm, the authors modelled the eccentric joint with a finite spring of length equal to the eccentricity of 50 mm. The spring
School of Civil Engineering Research Report No R896

Structural Modelling of Support Scaffold Systems

June 2009

had specific rotational stiffness, and was assumed to be axially stiff. The authors concluded that for large frames, unless torsion failure occurs, then the effects of joint eccentricity are insignificant. In the scaffold study by Gylltoft and Mroz [1] the braces were represented as truss members with pinned joints connected to the standards, and the connections between other members were modelled as short finite elements with nonlinear stiffness in all directions. To model the shores of the scaffold system, Peng et al. [15] applied rigid links with pinned supports at both ends, given that actual shores were connected loosely by nails at the top and bottom.

2. Full-Scale Subassembly Tests


2.1 Test Setup and Procedures A total of 18 support scaffold subassembly tests were conducted at the University of Sydney in 2006 [24] to study the behaviour and ultimate load-carrying capacities of such systems. In all tests, the formwork subassembly, also known as Cuplok scaffold system, was constructed as a grid frame of three-by-three bays with a constant nominal bay width of 1829 mm in both directions. The first fourteen tests were systems of three lifts with equal nominal lift height of 1.5 m. The systems featured different bracing arrangements (full, none, perimeter, core, and North-South direction only) and different jack extension heights (300 mm or 600 mm) at both top and bottom. In the last four subassembly tests, systems with 1 m and 2 m lift heights with full bracing were tested with the variation in the number of lifts from 2 to 4 and jack extension heights of 300 mm or 600 mm. All testing components were taken from stocks of used material. As a result, the testing materials showed some geometric imperfections representing those encountered in practice, particularly in regard to the out-of-straightness of the standards. Besides, the Cuplok joints showed signs of wear from frequent use and were therefore representative of joints used in practice in terms of joint stiffness and strength [24]. The standards, attached with Cuplok joints, were made from cold-formed circular steel tube grade 450 MPa with nominal outside diameter of 48.3 mm and thickness of 4 mm. The grade 350 MPa ledgers with end blades were of nominal outside diameter of 48.3 mm and thickness of 3.2 mm. The telescopic braces with hook ends were made of 48.3 mm x 4.0 mm outer tube and 38.2 mm x 3.2 mm inner tube with nominal yield stress of 400 MPa. The adjustable jacks were made of 36 mm diameter threaded steel rod of grade 430 MPa. The base plates were 180 mm x 180 mm x 10 mm in dimension with nominal yield stress of 250 MPa [24]. A test frame was constructed specially for the subassembly tests consisting of four loading beams at both the top and bottom running in the North-South direction. A total of sixteen hydraulic jacks attached to the top loading beams (four hydraulic jacks per loading beam) were used to load the (150 mm x 77 mm) timber bearers running in the East-West direction that applied loading to the top of each standard via (210 mm clear width) U-heads. Four sets of cross-bracing were installed to prevent sway of the top loading beams. Ball bearings were inserted between the jacks and timber bearers in the first test. However, this condition produced excessive lateral displacements and rotations of the adjustable jacks and was deemed non-representative of construction practice, where the bearers support a series of closely spaced secondary bearers running orthogonally to the main bearers. These secondary bearers elastically restrain displacements and rotations of the primary bearers. Consequently, the results of Test No. 1 were discarded and in subsequent tests, secondary bearers spaced approximately at 600 mm were attached at the top of primary bearers. Figure 6 illustrates a typical test frame showing primary and secondary bearers and bracings. The results of Test
School of Civil Engineering Research Report No R896

10

Structural Modelling of Support Scaffold Systems

June 2009

No. 7 were also discarded since the hydraulic jacks accidentally moved out of positions during the test.

Figure 6: Typical test frame from top view


School of Civil Engineering Research Report No R896

11

Structural Modelling of Support Scaffold Systems

June 2009

In all tests except Test No. 6, the loads were applied at an eccentricity of 25 mm in the NorthSouth direction to the top adjustable jacks (Figure 7(a)) along the second row in the EastWest direction while for the rest of the standards the loads were applied concentrically. In addition, the base plates in the row of eccentrically loaded standards were placed on 3 mm diameter circular steel rods at a nominal eccentricity of 15 mm, as shown in Figure 7(b), as per AS3610 [25]. The load eccentricities applied at the top and bottom of the system were arranged such that the standards were bent in single curvature. The loads were applied equally by hydraulic jacks on each standard through primary bearers, except in Test No. 14 where the loads applied to the corner, perimeter, and centre jacks were in the ratio of 1:2:4 respectively. The applied loads were recorded at each increment of loading until failure occurred, and theodolites were employed to measure initial geometric imperfections of all the standards before the test began and the displacements of six selected standards during loading [24].

Hydraulic jack Primary bearer

U-head Base plate

Unit: mm

3 mm steel rod (b)

(a)
Figure 7: Enlarged view of (a) top eccentricity and (b) bottom eccentricity 2.2 Test Configurations A summary of the test configurations which includes test number, test date, lift height, number of lifts, top and bottom jack extension length, position of spigot, bracing arrangement, type of loading, and loading eccentricity is presented in Table 1.
School of Civil Engineering Research Report No R896

12

Structural Modelling of Support Scaffold Systems

June 2009

Table 1: Summary of test configurations No. of lifts 3 3 3 3 3 3 3 3 3 3 3 3 3 3 2 3 4 4 Load eccentricity in 2nd row standards 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom 25 mm top & 15 mm bottom

Test

Date

Lift height 1.5 m 1.5 m 1.5 m 1.5 m 1.5 m 1.5 m 1.5 m 1.5 m 1.5 m 1.5 m 1.5 m 1.5 m 1.5 m 1.5 m 2m 1m 1m 1m

Jack extension 600 mm 600 mm 600 mm 600 mm 600 mm 600 mm 300 mm 300 mm 300 mm 300 mm 300 mm 300 mm 300 mm 300 mm 300 mm 600 mm 300 mm 300 mm

Spigot (lifts) 2nd & 3rd 2nd & 3rd 2nd & 3rd 2nd & 3rd 2nd & 3rd 2nd & 3rd 2nd & 3rd 2nd & 3rd 2nd & 3rd 2nd & 3rd 2nd & 3rd 2nd 2nd & 3rd 2nd & 3rd 2nd 3rd 3rd 3rd

Bracing

Loading

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18

17/1/06 31/1/06 8/2/06 13/2/06 16/2/06 22/2/06 3/3/06 10/3/06 17/3/06 24/3/06 30/3/06 7/4/06 13/4/06 3/5/06 15/5/06 19/5/06 7/6/06 30/6/06

full full full none

uniform uniform uniform uniform

perimeter uniform perimeter uniform full full none core full full N-S only core full full full full uniform uniform uniform uniform uniform uniform uniform 1:2:4 uniform uniform uniform uniform

A schematic of a typical test configuration is shown in Figure 8. The figure shows a typical 3lift support scaffold system with a lift height of 1.5 m in a full bracing arrangement being loaded by hydraulic jacks attached to the test frame. The figure also shows a temporary support deck for access to the test frame. The bracing arrangement is according to the labelling in the figure showing core and perimeter braces. Full bracing arrangement includes both core and perimeter braces. N-S bracing arrangement consists of braces running in the
School of Civil Engineering Research Report No R896

13

Structural Modelling of Support Scaffold Systems

June 2009

North-South direction. As an example, Figure 9 shows the actual Test No. 8 setup consisting of a 3-lift, 1.5 m lift height and 300 mm jack extension subassembly test with full bracing arrangement including top and bottom eccentricities in 2nd row.

Figure 8: Schematic of typical test configuration in plan and elevation view


School of Civil Engineering Research Report No R896

14

Structural Modelling of Support Scaffold Systems

June 2009

Figure 9: Test No. 8 setup 2.3 Test Results A summary of the test results consisting of test number, ultimate load from hydraulic jacks at 3 different locations (corner, perimeter, and centre), and observed failure mode is shown in Table 2. As noted earlier, the results of Tests No. 1 and 7 are unrepresentative, and thus not shown in the summary.
School of Civil Engineering Research Report No R896

15

Structural Modelling of Support Scaffold Systems

June 2009

Table 2: Summary of test results Ultimate load at corner jacks (kN) 87 91 50 60 60 130 65 70 120 119 70 40 105 100 140 150 Ultimate load at perimeter jacks (kN) 86 90 50 60 60 130 65 70 120 120 70 80 105 100 140 150 Ultimate load at centre jacks (kN) 89 91 50 60 60 130 65 70 120 120 70 160 105 100 140 150

Test

Observed failure mode

2 3 4 5 6 8 9 10 11 12 13 14 15 16 17 18

N-S sway mode, final failure of top and bottom jacks N-S sway mode, failure of top and bottom jacks, and spigot N-S sway mode, final failure of top and bottom jacks N-S sway of centre bay, final failure of top jacks N-S sway of centre bay, final failure of top jacks and spigot Some N-S sway, failure of standards and spigot at top lift N-S sway mode, final failure of top jacks and top standards N-S sway mode, failure of corner standards and spigots Some N-S sway, final failure of top spigots and standards Some N-S sway, failure of top spigots and corner standards Some E-W sway, final failure of perimeter spigots in 2nd lift Some N-S sway, failure of top spigots and centre standards Failure of corner spigot and top standard N-S sway mode, failure of jacks Bearer broke off before final failure of top jacks Some N-S sway, final failure of corner standard at top lift

The test results suggest that the failure modes are controlled by the jack extension length since when 600 mm top and bottom extensions are used the failure mode is North-South sway with final failure at the jacks. On the contrary, when 300 mm extensions are used, failure occurs mainly in the standards and spigots with only small sway displacements. Noticeably, the ultimate load decreases as the jack extension increases. The test results show that the bracing arrangement significantly influences the ultimate load of the system. Also, the higher lift height reduces the ultimate load. Moreover, the standards tend to fail at the top lift and around the perimeter region, especially at the corner where there is no bracing and only two ledgers are connected. Final failure occurred in spigots and jacks in most cases, as School of Civil Engineering 16 Research Report No R896

Structural Modelling of Support Scaffold Systems

June 2009

shown in Figure 10. Complete data on initial geometric imperfections and displacements as well as supplementary tests on components of Cuplok scaffold systems, including tensile coupon, stub column and bending tests are available in [24].

(a)

(b) Figure 10: Failure in (a) spigot and (b) jack

3. Finite Element Models


In this research, three-dimensional finite element models have been developed for analysing support scaffold systems. The analyses include geometric and material nonlinearities and are performed using the commercial finite element software package Strand7 [26]. The models present efficient and accurate methods for representing spigot joints, standard-to-ledger connections, base plate eccentricities, and load eccentricities. Also, initial geometric imperfections and material nonlinearity of all components of the system are incorporated in the models. In modelling the structural elements of the scaffold system, nonlinear beam elements, including contact, link, and connection elements, are used as described in sections 3.1 to 3.7. The models are compared with the subassembly tests [24] and calibrated against the ultimate loads and displacement responses in section 3.8. 3.1 Spigot Joints In the studied systems, the spigot joint consists of an insert made from a circular hollow steel tube with 38.2 mm outside diameter, 3.2 mm in thickness, and 300 mm in total length. The insert feeds into the abutting top and bottom standards to create a required lift height, as shown in Figure 2. The top standard can slide over the insert, which is fastened to the bottom standard by a fixed pin. The spigot modelling suggested by Enright et al. [19] is adopted, as shown in Figure 11.

School of Civil Engineering Research Report No R896

17

Structural Modelling of Support Scaffold Systems

June 2009

Load Pinned link Top Standard


150 mm

Pinned link Spigot Pinned link

Bottom Standard
150 mm

Figure 11: Schematic of spigot joint model The top and bottom standards are modelled as nonlinear beam element connected to the spigot via pinned connections. The spigot beam element is connected to the standards by three pinned stiff links capable of only transferring lateral forces from the standards to the spigot. As a result, the vertical force travels through the standards and only horizontal forces transfer to the spigot via the pinned links. When the standard bends under vertical load, the spigot is forced to bend because of the lateral forces acting oppositely at the top/bottom and centre of the spigot. The degree of bending of the spigot depends on the amount of initial geometric imperfection of the standard and vertical force. In three-dimensional analyses, the spigot model is arranged in the direction perpendicular to the primary bearers which is in the same direction as that of the load eccentricity. This arrangement is reasonable since it was observed from the subassembly tests [24] that the spigot joint tends to fail in the same direction as the load eccentricity. For simplicity, the spigot model is applied at mid height of the lift, even though the spigot is often located at little below or above mid height in actuality. 3.2 Semi-Rigid Standard-to-Ledger Connections The connections between standard and ledger in this research consist of a semi-rigid Cuploktype joint that can join up to four ledgers to the standard. The relation between the moment and rotation of the Cuplok connections is modelled by a tri-linear curve, as illustrated in Figure 12. The parameters that describe the tri-linear curve (k1, k2, k3, 1, 2, 3) were obtained from laboratory tests. Three different joint configurations were tested in bending about vertical and horizontal axes, i.e. the 4-way, 3-way and 2-way configurations, reflecting the number of ledgers connected at the joint. It was observed that the more ledgers connected, the less movement in the joint itself, and hence the greater stiffness. The average joint stiffness values for k1, k2, and k3 for different joint configurations and bending axes (Figure 13) are presented in Table 3. The average joint rotation values for 1, 2 and 3 are presented for different joint configurations and bending axes in Table 4. The connection element in Strand7 [26] is used to model the relation between moment and rotation. It requires that a multi-linear moment-rotation table is specified for bending about vertical and horizontal axes. The connection element is used to supply stiffness for any of the six degrees of freedom (axial, shear in 2 directions, and bending about 3 axes); in this case, only bending about vertical and horizontal axes is incorporated in the multi-linear table; the rest are assumed to be rigid.
School of Civil Engineering Research Report No R896

18

Structural Modelling of Support Scaffold Systems

June 2009

Table 3: Average Cuplok joint stiffness (kNm/rad) Bending about horizontal axis k1 k2 k3 80 102 5.3 75 87 5.1 70 77 4.6 Bending about vertical axis k1 k2 k3 15 7.5 0.8 14 7 1 7.5 5 1.5

Joint configuration 4-way 3-way 2-way

Table 4: Average rotation for Cuplok joints (rad) Bending about horizontal axis 1 2 3 0.014 0.036 0.16 0.012 0.036 0.16 0.007 0.036 0.16 Bending about vertical axis 1 2 3 0.02 0.04 0.1 0.02 0.04 0.1 0.02 0.04 0.1

Joint configuration 4-way 3-way 2-way

Moment k3

k2 k1 1 2 3 Rotation

Figure 12: Tri-linear moment-rotation for the Cuplok joints

Figure 13: Bending axes of the Cuplok joints 3.3 Brace Connections The braces are made of telescopic members with hooks at the ends. They are modelled using two rigidly connected elements with different cross-sections. Connection elements with only
School of Civil Engineering Research Report No R896

19

Structural Modelling of Support Scaffold Systems

June 2009

axial stiffness form the connection between the brace member elements and the ledgers. The axial spring stiffness is taken as 1.8 kN/mm as obtained from test calibrations on braced scaffold systems. The braces are offset 60 mm from the nodal points between the standards and ledgers because in actual construction the braces are connected to ledgers at about this distance away from the joints. 3.4 Base Plate Eccentricity The placement of the base plate of scaffold systems on an uneven or sloped ground can create eccentricity. The amount of base eccentricity depends largely on ground surface irregularities as shown in Figure 14(a). AS3610 [25] specifies an expected base eccentricity of no more than 40 mm or bp/4, whichever is less, where bp is the stiff portion of bearing of an end plate, as shown in Figure 14(a). For example, bp/4 is 17 mm for the scaffold system in the study which is less than 40 mm; therefore, the expected base eccentricity of the system is no more than 17 mm. The base eccentricity model proposed is illustrated in Figure 14(b), in which the base eccentricity is labelled as e. The standard and the base plate are modelled using nonlinear beam elements with their corresponding cross-sectional and material properties. A contact element is used to model a gap between the base plate and the ground. The contact element is set to provide stiffness only in compression, and only when the nodes to which it is connected come into contact, that is when the gap closes. The stiffness is specified as infinity, implying that when the load transfers from the standard to the base plate causing the far end of the base plate to rotate and touch the ground, the point of contact becomes infinitely stiff representing solid ground or other hard surface.

Load Standard Base plate


1 1

Standard

e
Uneven ground

Base plate Contact element

bp
(a) (b)

Figure 14: (a) base plate on uneven ground and (b) base eccentricity model 3.5 Load Eccentricity A load eccentricity can occur between the timber bearer and the U-Head since the bearer is not always positioned such that its centre line coincides with the centre line of the jack, even though in good construction practice the U-head is twisted against the bearer so as to reduce the amount of eccentricity. In order to model the eccentricity, a rigid link with length equal to the load eccentricity is connected to the top of the jack in the direction perpendicular to the bearer, and the vertical point load is applied at the far end of the link. The rigid link behaves as a short, stiff cantilever that introduces vertical force and additional moment into the jack. 3.6 Geometric Imperfections Scaffold systems are generally slender and sensitive to stability effects; therefore, initial geometric imperfections producing member P- and frame P- effects (Figure 15) must be considered in the analysis model. A common approach to incorporate geometric
School of Civil Engineering Research Report No R896

20

Structural Modelling of Support Scaffold Systems

June 2009

imperfections is to scale one or more critical elastic buckling modes and apply the scaled displacements to the perfect geometry. Nevertheless, several issues remain unanswered in this method. For instance, how many buckling modes should be included and what scaling factors should be applied? One alternative is to use the maximum allowable imperfection values from available structural codes. For example, the Australian steel design standard [7] specifies a maximum out-of-straightness of L/1000, where L is the member length. Nonetheless, this method often produces conservative predictions of the strength of systems. Another approach is to apply notional horizontal forces, but some doubt remains as to the magnitude of force to use. Alternatively, for the purpose of calibrating the analysis model, in this research the magnitudes of the member out-of-straightness and the frame out-of-plumb are implemented directly at the nodes of the finite element models from acquired initial imperfection measurements taken as part of the subassembly tests [24]. The member out-of-straightness () is applied at mid height of the standard in each scaffold lift and the frame out-of-plumb () is applied at each ledger-standard connection point and at the U-head at the top of the scaffold. From the measurements, the average out-of-straightness of the standards is Lh/820 mm for the lifts with spigot joint and Lh/1700 mm for the lifts without spigot joint, where Lh is the lift height, and the average out-of-plumb of the frames is H/470 mm, where H is the total height. Full details of the initial geometric imperfections of the subassembly tests are available in [24]. For other studies on the scaffold systems, real data on initial geometric imperfections is procured from construction sites around the Sydney area.

Figure 15: P- and P- effects 3.7 Geometric and Material Nonlinearities In geometric nonlinear analysis, an accurate determination of the displacements is attained which is particularly important in slender structures such as scaffold systems. In the present beam element based analysis, the deformed geometry is used to establish the equilibrium equation and the elements local reference system is updated at each load increment to capture the load deflection characteristics. In material nonlinear analysis, the nonlinear relationship between stress and strain is applied. The stress-strain relations for the scaffold components used in the models are based on the Ramberg-Osgood expression [27, 28] fitted to experimental data obtained from supplementary tests on components of the subassembly tests [24]. Figures 16 to 21 show the Ramberg-Osgood stress-strain curves used in the material modelling of the experimental standard, ledger, jack, base plate, brace, and spigot respectively. Since there is no stress-strain data for the ledger, brace and spigot, the stress-strain relations for these components are obtained by scaling the Ramberg-Osgood stress-strain relation used for the standards to their
School of Civil Engineering Research Report No R896

21

Structural Modelling of Support Scaffold Systems

June 2009

nominal yield stress. However, the ledger, brace and spigot insert are expected to be loaded only in elastic range, and therefore, the material nonlinearity effects of these components are negligible. The Ramberg-Osgood parameters (E0, 0.2, n) for each scaffold component are summarised in Table 5. In the table, E0 is the initial Youngs modulus, 0.2 is the 0.2% proof stress as the equivalent yield stress, and n is a parameter which determines the sharpness of the knee of the stress-strain curve. Table 5: Ramberg-Osgood parameters for scaffold components Component E0 (GPa) 0.2 (MPa) Standard Ledger Jack Base plate Brace Spigot 200 200 200 200 200 200 530 380 495 260 430 430 n 38.2 38.2 16.0 25.0 38.2 38.2

The Ramberg-Osgood stress-strain relations are applied to the beam elements of each scaffold component. As the beam cross-section and length are subdivided, sampling points on the cross-section and integration points along the length of the beam are utilised to numerically integrate the stiffness characteristics of the beam. As a result of this section and length-wise integration, the propagation of yielding through the cross-section and along the beam element can be included. The axial and bending stiffness are coupled as the neutral axis on the yielded beam shifts [26]. This method is referred to as plastic-zone analysis [8].
6.00E+08 5.00E+08 4.00E+08 Stress (Pa) 3.00E+08 2.00E+08 1.00E+08 0.00E+00 0.0000 0.0050 0.0100 Strain 0.0150 0.0200 0.0250

Ramberg-Osgood

Test result

Figure 16: Stress-strain curve for standard


School of Civil Engineering Research Report No R896

22

Structural Modelling of Support Scaffold Systems

June 2009

5.00E+08

4.00E+08 Stress (Pa)

3.00E+08

2.00E+08

1.00E+08

0.00E+00 0.0000 0.0050 0.0100 Strain Ramberg-Osgood 0.0150 0.0200 0.0250

Figure 17: Stress-strain curve for ledger

7.00E+08 6.00E+08 5.00E+08 Stress (Pa) 4.00E+08 3.00E+08 2.00E+08 1.00E+08 0.00E+00 0.0000 0.0050 0.0100 Strain Ramberg-Osgood Test result 0.0150 0.0200 0.0250

Figure 18: Stress-strain curve for jack

School of Civil Engineering Research Report No R896

23

Structural Modelling of Support Scaffold Systems

June 2009

4.00E+08

3.00E+08 Stress (Pa)

2.00E+08

1.00E+08

0.00E+00 0.0000 0.0050 0.0100 Strain Ramberg-Osgood Test result 0.0150 0.0200 0.0250

Figure 19: Stress-strain curve for base plate

5.00E+08

4.00E+08 Stress (Pa)

3.00E+08

2.00E+08

1.00E+08

0.00E+00 0.0000 0.0050 0.0100 Strain Ramberg-Osgood 0.0150 0.0200 0.0250

Figure 20: Stress-strain curve for brace

School of Civil Engineering Research Report No R896

24

Structural Modelling of Support Scaffold Systems

June 2009

5.00E+08

4.00E+08 Stress (Pa)

3.00E+08

2.00E+08

1.00E+08

0.00E+00 0.0000 0.0050 0.0100 Strain Ramberg-Osgood 0.0150 0.0200 0.0250

Figure 21: Stress-strain curve for spigot 3.8 Calibrations The commercial software package Strand7 [26] was used to create a finite element model for each of the full-scale subassembly tests using the actual frame dimensions and measured values of imperfections. The mean of the measured dimensions of components were used for cross-sectional properties in the finite element models. As an example, Figure 22 shows the finite element model for Test No. 3 of the subassembly tests [24]. The ultimate loads and displacements obtained from the nonlinear analyses accounting for both material and geometric nonlinearities were calibrated against failure loads and load-deflection responses obtained from the full-scale subassembly tests [24]. The calibrations were achieved by changing the stiffness of the elastic restraints applied at the U-head and base plate, as well as the axial spring stiffness of the brace connections; the latter was changed after the calibrations were performed on unbraced systems for the top and bottom rotational stiffness. Table 6 shows the results of the stiffness parameters obtained from the calibrations. In the table, K represents translational stiffness with subscript showing its direction corresponding to Figure 22, and R represents rotational stiffness with subscript showing the axis of bending according to Figure 22. It should be noted that the top rotational stiffness about the x-axis is assumed to be rigid, corresponding to the negligible strong axis bending of the bearer. However, the y-axis bending stiffness is taken as 40 kNm/rad since bending about this axis occurs during failure as observed in the tests [24]. The bottom rotational stiffness about the x and y axes is calibrated as 100 kNm/rad. The bottom rotational stiffness is applied to all uprights except the uprights with bottom eccentricity for which base plate modelling is applied. The translational stiffness at the base is taken as rigid in all directions. At the top, the translational stiffness is assumed to be rigid in the x and y directions, but 0 in the z direction. The brace end connections have an axial stiffness of 1.8 kN/mm capable of transferring only axial forces to the ledgers.
School of Civil Engineering Research Report No R896

25

Structural Modelling of Support Scaffold Systems

June 2009

Table 7 shows the calibration results for the failure loads and their statistics. The average of the ratio of test failure load to ultimate load from advanced analysis is 1.014 with standard deviation (STD) and coefficient of variation (COV) of 0.0980 and 0.0966, respectively. Figures 23 to 37 compare the finite element analysis results with the experimental loaddeflection responses at certain point of the frame, as indicated for each test in the titles of the figures. Table 6: Parametric calibration results
Bottom boundary conditions Kx (kN/mm) Ky (kN/mm) Kz (kN/mm) Rx (kNmm/rad) Ry (kNmm/rad) Rigid Rigid Rigid 100,000 100,000 Top boundary conditions Kx (kN/mm) Ky (kN/mm) Kz (kN/mm) Rx (kNmm/rad) Ry (kNmm/rad) Rigid Rigid 0 Rigid 40,000 Brace end connections Axial stiffness (kN/mm) 1.8

Rz (kNmm/rad) 0

Rz (kNmm/rad) 0

Table 7: Load calibration results


Ultimate load from Test failure advanced load (kN) analysis (kN) 96 89 91 91 45 50 60 60 66 60 138 130 50 65 64 70 127 120 129 120 68 70 160 160 105 105 100 100 147 150 Average STD COV

Test 2 3 4 5 6 8 9 10 11 12 13 14 15 16 18

Test load / Advanced analysis result 0.927 1.000 1.111 1.000 0.909 0.942 1.300 1.094 0.945 0.930 1.029 1.000 1.000 1.000 1.020 1.014 0.0980 0.0966

School of Civil Engineering Research Report No R896

26

Structural Modelling of Support Scaffold Systems

June 2009

Figure 22: Finite element model of Test No. 3 showing axes

120 100 80 Load (kN) 60 40 20 0 0 5 10 15 20 Deflection (mm) Test result FE result 25 30 35 40

Figure 23: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 2

School of Civil Engineering Research Report No R896

27

Structural Modelling of Support Scaffold Systems

June 2009

100 90 80 70 Load (kN) 60 50 40 30 20 10 0 0 5 10 15 20 Deflection (mm) Test result FE result 25 30 35 40

Figure 24: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 3

50 45 40 35 Load (kN) 30 25 20 15 10 5 0 0 5 10 15 Deflection (mm) Test result FE result 20 25 30

Figure 25: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 4

School of Civil Engineering Research Report No R896

28

Structural Modelling of Support Scaffold Systems

June 2009

70 60 50 Load (kN) 40 30 20 10 0 0 5 10 15 20 25 30 35 Deflection (mm) Test result FE result

Figure 26: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 5

70 60 50 Load (kN) 40 30 20 10 0 0 2 4 6 Deflection (mm) Test result FE result 8 10 12

Figure 27: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 6

School of Civil Engineering Research Report No R896

29

Structural Modelling of Support Scaffold Systems

June 2009

160 140 120 Load (kN) 100 80 60 40 20 0 0 5 10 15 Deflection (mm) Test result FE result 20 25 30

Figure 28: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 8

60 50 40 Load (kN) 30 20 10 0 0 1 2 3 4 5 6 7 8 9 Deflection (mm) Test result FE result

Figure 29: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 9

School of Civil Engineering Research Report No R896

30

Structural Modelling of Support Scaffold Systems

June 2009

70 60 50 Load (kN) 40 30 20 10 0 0 0.5 1 1.5 Deflection (mm) Test result FE result 2 2.5 3

Figure 30: Calibration of load-deflection responses at mid-height of the standard of the 3rd lift of the 2nd row of the frame for Test No. 10

140 120 100 Load (kN) 80 60 40 20 0 0 2 4 6 8 Deflection (mm) Test result FE result 10 12 14 16

Figure 31: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 11

School of Civil Engineering Research Report No R896

31

Structural Modelling of Support Scaffold Systems

June 2009

140 120 100 Load (kN) 80 60 40 20 0 0 2 4 6 8 10 12 14 16 18 Deflection (mm) Test result FE result

Figure 32: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 12

80 70 60 Load (kN) 50 40 30 20 10 0 0 0.5 1 1.5 2 2.5 Deflection (mm) Test result FE result 3 3.5 4 4.5 5

Figure 33: Calibration of load-deflection responses at the 2nd lift of the 1st row of the frame for Test No. 13

School of Civil Engineering Research Report No R896

32

Structural Modelling of Support Scaffold Systems

June 2009

180 160 140 120 Load (kN) 100 80 60 40 20 0 0 2 4 6 8 10 12 14 16 18 Deflection (mm) Test result FE result

Figure 34: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 14

120 100 80 Load (kN) 60 40 20 0 0 2 4 6 8 Deflection (mm) Test result FE result 10 12 14 16

Figure 35: Calibration of load-deflection responses at mid-height of the standard in the 2nd lift of the 2nd row of the frame for Test No. 15

School of Civil Engineering Research Report No R896

33

Structural Modelling of Support Scaffold Systems

June 2009

120 100 80 Load (kN) 60 40 20 0 0 5 10 15 Deflection (mm) Test result FE result 20 25 30

Figure 36: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 16

160 140 120 Load (kN) 100 80 60 40 20 0 0 5 10 15 20 25 30 35 Deflection (mm) Test result FE result

Figure 37: Calibration of load-deflection responses at mid-height of the standard in the 3rd lift of the 2nd row of the frame for Test No. 18

School of Civil Engineering Research Report No R896

34

Structural Modelling of Support Scaffold Systems

June 2009

4. Discussion
The calibrations show that advanced analysis using geometric and material nonlinear finite element models gives very good predictions of the ultimate loads of the systems. Most of the predictions are within 10% of the actual failure loads. In fact, the average of the ratios between failure test load and predicted ultimate load is very close to 1 (1.014) with a relatively small COV of 0.0966. In addition, advanced analysis gives good results in predicting deformation responses of support scaffold systems. The finite element analysis results of the load-deflection responses fit the test results [24] reasonably closely with most of the values within 20% of one another. It can be noticed that in some tests there are no deflection values available for the failure load. Also, Test No. 4 only provides three measured deflections during loading, and has been ignored in the comparison. Two distinct failure modes are observed from the advanced analysis, as one exhibiting an Sshape member buckle (Figure 38) and the other a lateral frame buckle with large lateral displacements at the top story (Figure 39). The failure modes are noticed to be sensitive to the jack extension length, where 600 mm jack extension produces lateral frame buckling with main failure in the jacks and 300 mm jack extension produces S-shape buckling of the standards with predominant failure deformations in the spigots. These failure modes were also observed in the tests [24] suggesting that advanced analysis is capable of accurately predicting the behaviour and failure mode of support scaffold systems.

Figure 38: S-shape member buckling

School of Civil Engineering Research Report No R896

35

Structural Modelling of Support Scaffold Systems

June 2009

Figure 39: Lateral frame buckling

5. Conclusions
In this paper, nonlinear finite element analysis models for support scaffold systems have been developed. Models for various components of the systems including spigot joints, semi-rigid upright-to-beam connections and base plate eccentricities are proposed. Calibrations of these models to the full-scale subassembly tests [24] consisting of three-by-three bay formwork systems with the combinations of different numbers of lifts, jack extension, and lift height are achieved by adjusting the top and bottom boundary conditions as well as the brace connection stiffness. The ultimate loads obtained from advanced analysis are in close agreement with the failure loads of the tests; moreover, comparisons of load-deflection responses also show close agreement, demonstrating that advanced analysis is able to accurately predict the behaviour and strength of highly complex support scaffold systems. The development of a design methodology for support scaffold systems based on advanced analysis is in progress at the University of Sydney.

Acknowledgement
The authors would like to thank Boral Formwork & Scaffolding Pty Ltd for providing subassembly test data and support of this research project.

References
[1] Gylltoft K, Mroz K. Load carrying capacity of scaffolds. Structural Engineering International 1995;1:37-42. [2] Peng JL, Chan SL, Wu CL. Effects of geometrical shape and incremental loads on scaffold systems. Journal of Constructional Steel Research 2007;63(4):448-459. [3] Prabhakaran U, Godley MHR, Beale RG. Three-dimensional second order analysis of scaffolds with semi-rigid connections. Welding in the World 2006;50(SPEC ISS):187-194. [4] Yu WK, Chung KF, Chan SL. Structural instability of multi-storey door-type modular steel scaffolds. Engineering Structures 2004;26(7):867-881.
School of Civil Engineering Research Report No R896

36

Structural Modelling of Support Scaffold Systems

June 2009

[5] Chu AYT, Chan SL, Chung KF. Stability of modular steel scaffolding systems theory and verification. Proceedings of International Conference Advances in Building Technology Hong Kong, 2002, pp. 621-628. [6] Weesner LB, Jones HL. Experimental and analytical capacity of frame scaffolding. Engineering Structures 2001;23(6):592-599. [7] Standard Australia. AS 4100: Steel structures. 1998. [8] Clarke MJ, Bridge RQ, Hancock GJ, Trahair NS. Advanced analysis of steel building frames. Journal of Constructional Steel Research 1992;23(1-3):1-29. [9] ANSYS Release 5.5. ANSYS basic analysis procedures guide. Swanson Analysis Systems, 1998. [10] LUSAS. User manual version 12.2. FEA Ltd, UK, 1998. [11] NAF-NIDA. Software for nonlinear integrated design and analysis version 3. User's manual. Department of Civil and Structural Engineering, Hong Kong Polytechnic University, 2001. [12] Milojkovic B, Beale RG, Godley MHR. Determination of the factors of safety of standard scaffold structures. Proceedings of International Conference Advances in Steel Structures, Vol. 1 UK, 2002, pp. 303-310. [13] Godley MHR, Beale RG. Sway stiffness of scaffold structures. Structural Engineer 1997;75(1):4-12. [14] Huang YL, Chen HJ, Rosowsky DV, Kao YG. Load-carrying capacities and failure modes of scaffold-shoring systems, Part I: Modeling and experiments. Structural Engineering and Mechanics 2000;10(1):53-66. [15] Peng JL, Pan ADE, Chen WF, Yen T, Chan SL. Structural modeling and analysis of modular falsework systems. Journal of Structural Engineering 1997;123(9):1245-1251. [16] Chan SL, Huang HY, Fang LX. Advanced analysis of imperfect portal frames with semirigid base connections. Journal of Engineering Mechanics 2005;131(6):633-640. [17] Yu WK, Chung KF. Prediction on load carrying capacities of multi-storey door-type modular steel scaffolds. Steel & Composite Structures 2004;4(6):471-487. [18] Godley MHR, Beale RG. Analysis of large proprietary access scaffold structures. Proceedings of the Institution of Civil Engineers: Structures and Buildings, Vol. 146 UK, 2001, pp. 31-39. [19] Enright J, Harriss R, Hancock GJ. Structural stability of braced scaffolding and formwork with spigot joints. Proceedings of the Fifteenth International Specialty Conference on Cold-Formed Steel Structures, Vol. 1 St. Louis, Missouri, USA, 2000, pp. 357-376. [20] Yu WK. An investigation into structural behaviour of modular steel scaffolds. Steel & Composite Structures 2004;4(3):211-226. [21] Vaux S, Wong C, Hancock G. Sway stability of steel scaffolding and formwork systems. Proceedings of the Third International Conference on Advances in Steel Structures, Vol. 1 Hong Kong, 2002, pp. 311-319. [22] Harung HS, Lightfoot E, Duggan DM. The strength of scaffold towers under vertical loading. Structural Engineer 1975;53(1):23-30. [23] Milojkovic B., Beale R.G., M.H.R. G. Modelling scaffold connections. Proceedings of the 4th ACME UK Annual Conference 1996:85-88. [24] CASE. Tests of formwork subassemblies and components. Investigation Report No. S1499. Centre for Advanced Structural Engineering, School of Civil Engineering, University of Sydney, 2006. [25] Standard Australia. AS 3610: Formwork for concrete. 1995. [26] Strand7. Strand7 release 2.4 manual. Strand7 Pty Ltd, 2009. [27] Ramberg W, Osgood WR. Determination of stressstrain curves by three parameters. Technical Note No. 503. National Advisory Committee on Aeronautics (NACA). 1941. [28] Rasmussen KJR. Full-range stress-strain curves for stainless steel alloys. Journal of Constructional Steel Research 2003;59(1):47-61.
School of Civil Engineering Research Report No R896

37

S-ar putea să vă placă și