Vous êtes sur la page 1sur 13

Engineering Structures 32 (2010) 12231235

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Nonlinear dynamic simulations of progressive collapse for a multistory building


Leslaw Kwasniewski
Warsaw University of Technology, Al. Armii Ludowej 16, 00-637 Warszawa, Poland

article

info

abstract
The paper presents a case study of progressive collapse analysis of a selected multistory building. The subject of the numerical study is an existing 8-story steel framed structure built for fire tests in the Cardington Large Building Test Facility, UK. The problem is investigated using nonlinear dynamic finite element simulations carried out following the GSA guidelines. The paper focuses on model development for global models subject to increasing vertical loading and notional column removal. Taking advantage of parallel processing on multiprocessor computers, a detailed 3D model with large number of finite elements has been developed for the entire structure. The objective of the presented feasibility study is to identify modeling parameters affecting the final result (potential of progressive collapse) and propose a hierarchical verification and validation program for reducing outcome uncertainties. 2009 Elsevier Ltd. All rights reserved.

Article history: Received 23 August 2009 Received in revised form 9 November 2009 Accepted 22 December 2009 Available online 13 January 2010 Keywords: Progressive collapse Robustness Computer simulation Finite element model Cardington test Nonlinear transient dynamic analysis Verification Validation

1. Introduction After several disastrous building collapses, concepts such as progressive collapse and robustness of structures have been reflected in many research papers and resulted in new codes and guidelines available in Europe [1] and in the United States: [2,3]. The collapse of an entire structure or an essential part of it that is disproportionately large compared to the initiating local damage is considered a progressive collapse. In addition to the design guidelines, the mentioned standards provide provisions for the progressive collapse analysis of newly designed and existing structures. The main objective of such analysis is the assessment of the potential for progressive collapse. The behavior of the structure is analyzed in terms of the alternate load paths, tie forces, connection redundancy and resilience, and catenary or compressive arching actions of the structural members [3]. The studies require consideration of many structural features not included in the original design, such as inelastic material properties and limit values, damage criteria for structural members, and large deformations. Numerous potential causes can be considered for initial structural damage such as gas explosions, terrorist attacks, faulty construction, foundation failure, or accidental vehicle impacts. There are also possible numerous configurations for abnormal loading resulting from such disastrous accidents.

The concept for progressive collapse analysis should determine provisions for three aspects: loading configurations including abnormal loads, global failure criteria quantitatively defining the collapse phenomenon, and adequate analysis methods. In the most common threat independent methodology, the alternate load path method, abnormal primary loading is conceptually represented by the notional removal of major bearing structural elements [4]. Then the remaining structure is analyzed to determine if the initiated damage can propagate. For typical framed structures, the US General Services Administration (GSA) [3] recommends three considerations for the external columns (middle of the short side, middle of the long side, and the corner of a building) and one for the internal, all located at the ground-floor level. The DoD [2] applies the method at each floor level in a multistory building. At the moment of the structural element removal, the analyzed structure is subject to the configuration of design dead and live loads, with proportionality factors depending on the selected analysis method [2]. Increased factors are used in static analyses to account for actual dynamic loads generated by the structural motions caused by the sudden failure of an element (i.e., secondary loads) [4]. The GSA Guidelines [3] define the allowable extent of collapse as the smaller of the following two areas:

the area limited to structural bays directly associated with the


instantaneously removed vertical member and located directly above the removed member, 167 m2 (1800 ft2 ) at the floor level directly above the instantaneously removed vertical member.

Tel.: +48 22 660 6381. E-mail address: l.kwasniewski@il.pw.edu.pl.

0141-0296/$ see front matter 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.engstruct.2009.12.048

1224

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

GSA [3] introduced a flow-chart procedure for determining if a building under consideration can be exempt from detailed consideration for progressive collapse. For non-exempt structures, finite element analyses on different complexity levels are proposed. One can choose among linear or nonlinear, static or dynamic (time history), and between 2- and 3-dimensional (2D and 3D) analysis techniques [5]. GSA [3] limits the applicability of linear elastic static analysis procedures to buildings with a maximum of 10 stories above the ground. The nonlinear dynamic procedures are considered the most sophisticated and accurate structural assessment techniques. However, due to the complexity and inherent challenges, these procedures have been used less frequently for progressive analysis [3]. The main difficulties are related to the numerical convergence problems, material models capturing inelastic properties and damage, modeling component disintegration caused by failure, requirements for mesh resolution capturing local effects, and the size of the FE models for large structures (e.g., multistory buildings). The paper presents a case study of progressive collapse analysis of a selected multistory building. Different loading configurations are analyzed through nonlinear dynamic simulations using the commercial program LS-DYNA with explicit time integration. Taking advantage of parallel processing on multiprocessor computers, a detailed model has been developed for the actual steelframed building, which was originally constructed for full-scale fire tests. The objective of the presented feasibility study is to identify modeling parameters affecting the final result (potential of progressive collapse) and propose a verification and validation program for reducing outcome uncertainties. 2. Recent studies on modeling progressive collapse Most of the published progressive collapse analyses for entire buildings or their components are based on the alternate load path method with column removal. The previously mentioned standards [2] and [3] are usually adopted regarding load configurations and quantification of collapse. Differences are encountered in the numerical technique applied to predict structural behavior. Marjanishvili [4] evaluated, in a general manner, four successively more sophisticated analysis procedures for estimating the progressive collapse hazard: linear-elastic static, nonlinear static, linear-elastic dynamic, and nonlinear dynamic. The nonlinear time history (dynamic) analysis is recognized as giving the most realistic results, but at the same time, due to its high complexity, it is prone to incorrect assumptions and modeling errors. An example of linear static analysis for a cold-formed steel framed structure, the army barracks, is presented in [6]. In this step-by-step procedure, the members that exceed the so-called Demand-Capacity Ratios (DCR) [3] are removed, linear analysis is conducted for the changed configuration, and the new DCR values are calculated. The procedure is repeated until all the members left have DCR values smaller than the limit magnitudes, and then the extent of the structural damage is evaluated. Khandelwal et al. [7] analyzed the progressive collapse potential of seismically designed steel-braced frames, using explicit transient dynamic simulations. The study used the alternate path method on previously designed 10-story prototype buildings. The structural response was predicted using calibrated 2-dimensional macromodels built as a combination of beamcolumn and discrete spring finite elements. Izzuddin et al. in two papers [8] and [9] presented a design-oriented methodology for progressive collapse assessment of multistory buildings. The proposed assessment framework consists of three stages: nonlinear static response of the damaged structure under gravity loading, a simplified dynamic assessment to establish pseudo static curves, and ductility assessment of the

connections. The second paper details the application of the new approach to progressive collapse assessment of real steel-framed composite multistory buildings. In two papers, [10,11] presented an approach for a simulation of the inverse problem optimization of the structural collapse initiated using controlled explosives. The authors multilevel methodology, oriented toward uncertainty analysis, is based on multibody models (MSC.Adams) accompanied by a priori finite element analyses (FEAP) and by transient finite element calculations (LS-DYNA) performed on the computational cluster. The simplified multibody simulations are then implemented for fuzzy analysis. Lynn and Isobe [12] presented structural collapse analysis of a framed structure subject to impact loads. A high-rise framed structure impacted by a small aircraft is analyzed, using an FE model built of beam elements with the modified adaptively shifted integration (ASI) technique applied for through sectional integration. In this technique, formation of a plastic hinge in an element is realized by shifting the numerical integration point. The beam element formulation is also applied for progressive collapse dynamic analysis in [13]. Through use of damagedependent constitutive relationships, the developed beamcolumn element accounts for the interaction of the axial force and bending moment, including strength and stiffness degradation. The paper presents an example analysis of the planar frame structure subject to rapid column removal. Kim et al. [14] applied the FE open source code OpenSees for development of an integrated system of progressive collapse analysis. The system is supposed to allow user-friendly modeling, nonlinear dynamic analysis, graphic simulation, and evaluation of damage. The numerical examples presented are limited to the 2- and 3-story framed structures represented by 2-dimensional planar frames. Analyses of the collapse with the aid of the 2D FE model built of structural 1-dimensional (1D) (beam and truss) elements can also be found in [15]. The paper investigates the stability of the Twin Towers of the World Trade Center, New York, for different fire scenarios. A combined finite-discrete element model for failure and collapse of structural systems is presented in [16]. The paper presents development of a two-node finite element with numerical integration, allowing the capture of the non-linear behavior of both concrete and reinforcement. The numerical results are confined to simply supported reinforced concrete beams subject to four-point bending. Two nonlinear analysis methods are presented in [17] for a simplified evaluation of the progressive collapse potential in weldedsteel moment frames. First, the load-resisting mechanism of the column-removed double-span beams is numerically investigated based on shell FE submodels. Then the developed model is applied in the energy-based nonlinear static progressive collapse analysis based on the assumption of sufficiently strong and ductile connections. Mohamed [18] considered a case study for the progressive collapse analysis of a reinforced concrete building using the Alternate Path (AP) method according to the guidelines [2]. Presented numerical case studies based on the linear static analysis showed the importance of incorporating 3-dimensional effects, especially at the part of the structure where a column is notionally removed. A review of recently published numerical studies of progressive collapse behavior shows some clear tendencies. In most of the work, commercial nonlinear FE programs are implemented, such as: ABAQUS [15,17], ADAPTIC [8,9], FEAP [10,11], LS-DYNA [7,11], SAP2000 [4,6]. Beam element models dominate, and most of the considerations are confined to 2D subsystems. Numerous

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

1225

Fig. 1. General floor layout of the Cardington steel frame building, [20,24], with locations of removed columns.

simplifications applied in the models are justified by the required limitation of the computational time and resources. In more than half of the mentioned works, the multilevel strategy is applied, where the structure is analyzed first on the subsystem or component level before global analysis is performed on a simplified global model. 3. Description of the selected structure The design of an existing 8-story (33 m) building, built inside a former airship hanger at the Cardington Large Building Test Facility (LBTF) in the UK, was selected as the source of information for the subject of the feasibility study presented in this paper. This building, constructed for large-scale fire tests, was selected as a well-documented, representative example of a modern multistory office building designed to meet the most up-to-date British and European standards [19]. Most of the design data can be found in public domain reports [2022] and numerous papers dealing with global analysis of fire tests conducted at this facility; e.g., [19,23]. The test building is a steel-framed structure with a light concrete slab cast onto profiled steel decking and supported on a network of secondary and primary steel beams. The composite slabs include anti-crack meshes of 142 mm2 /m (T6@200 mm) in both directions, placed 15 mm above the steel deck [19]. Composite action is achieved by protrusions in trapezoidal steel decking bonded to the concrete. According to the original design, the frame structure is supposed to be braced by three stiff cores: a central lift shaft and two staircases at either end of the building. In the actual test building, the cores are represented by crossbar bracing (see Fig. 1). The typical floor plan (45 m 21 m) of the building, shown in Fig. 1, consists of five bays (5 9 m) in the longitudinal direction and three bays (6, 9, 6 m) in the transverse direction. The building was designed for a dead load of 3.65 kN/m2 and a live load of 3.5 kN/m2 [19]. To reduce fabrication and erection costs, only four beam sections (254 UB trimmers, 305 UB ribs, and 356 UB and 610 UB spine members) and three column sections (305 UC 198 and 118 kg/m and 254 UC 89 kg/m) were used for the steel frame [20]. The structure, designed primarily for gravity loads, consists of nominally simple (designed to transmit vertical shear only) joints with flexible flush 8-mm plates for beam-to-column connections and fin 10-mm plates for beam-to-beam connections, all with M20 bolts. The material properties of the steel and steel connectors (at ambient temperature) are presented in Tables 1 and 2 [23].

4. FE model development 4.1. General assumptions The objective of the presented feasibility study is to develop a detailed FE model appropriate for large-scale analysis of the selected structure and to take advantage of the computational capabilities of multiprocessor computers. Nonlinear transient dynamic analysis has been chosen as the method giving the most realistic results [4]. Computer simulations imitating the structural behavior were based on the commercial FE program LSDYNA applying the central difference method for the explicit time integration [25]. The chosen program is highly parallelized and perfectly suited to run on multiprocessor clusters. To effectively model the selected building (or its large portions), several simplifying assumptions had to be made. Only the skeleton of the building (stage 3 of the construction [22]), consisting of the steel framework and concrete slabs, has been modeled. The walls built of hollow blocks in stage 4 of the construction are not represented in the present study as a structural element in the FE models. The general floor plan presented in Fig. 1, and consistent with the floor arrangement in the actual building, was applied to 3 to 7 floors in the FE model. The differences in the plan arrangements (e.g., entrance opening) in the ground floor and first floor in the actual structure are included in the FE model. The bracing effect of the concrete cores (the central lift shaft and two staircases) is represented in the FE model by truss elements standing for the crossbars applied in the actual test building [24]. There are also some other variations between the FE model and the actual building. Some are due to missing information about construction details, and some are a consequence of differences between the actual structure and the design. For example, mistakenly placing the anti-crack reinforcement meshes directly on top of the steel decking caused cracks along all the primary steel beams [19]. Thus, the FE model, subject to the feasibility study at the current stage, is not a close replica of the actual 8-story steelframed building at the Cardington LBTF but should be considered a numerical representation of a realistic multistory structure. The complete FE model of the entire structure consisting of 1.08 million of finite elements is shown in Fig. 2. 4.2. Loading and boundary conditions All columns are fixed at their bases with the appropriate displacement and rotation conditions. The ground is modeled as a socalled rigid wall [25] the planar geometric boundary condition

1226

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

Table 1 Material properties for structural steel applied in the FE models. Property Ratios Magnitudes ASTM A36 [27] Yield stress Yield to ultimate stress ratio Yield strain End of yield Hardening strain Strain hardening modulus Failure (total) strain fy (MPa) fy /fu y = fy /E 331 0.71 0.00158 10 25 73 136 Magnitudes applied in the FE model S275 303 0.65 0.00144 6 [28] 28 (b = 4% [28]) 78 (ESH = 2.7 GPa [28]) 207 [29] S355 469 0.73 0.00189 6 [28] 21 (b = 4% [28]) 78 (ESH = 2.7 GPa [28]) 140 [29]

st /y b /y E /ESH f /y

by GSA [3]. Column removal is applied using a special option (*LOAD_REMOVE_PART) that allows for gradual reduction of stresses prior to the element deletion to prevent shock effects. In the computer simulations, the reduction of stresses at selected elements starts at 1 s, and the whole column is removed at t = 2 s (see Fig. 3). 4.3. Structural steel and bolts In steel and composite, steelconcrete, frame structures, being in the inelastic range, deformation is concentrated in plastic hinges and zones. The performance of the plastic hinges, usually localized at the joints, depends strongly on the inelastic properties of the steel components. The critical parameters are yield stress fy (or proof stress), tangent modulus for strain hardening ESH , ultimate stress fu , and failure strain f (or elongation after failure). If only nominal properties are known for the specific structural steel grade, simplified material models have to be applied, such as the elastic ideally plastic or bilinear elastic plastic model with hardening, recommended by Eurocode 3 [26]. In the current study, a more detailed and accurate general model for structural steel proposed by Galambos [27], and presented in Fig. 4, was applied. A survey of tensile (mill) tests in [28] shows that although there is a clear link between the yield point and material thickness (e.g., flange thickness), the strain hardening behavior is independent of the material thickness and steel grade, allowing a generalization of the stressstrain relationship for hot rolled steel. Magnitudes of the ratios defining the characteristic points in the diagram shown in Fig. 4 for steel grades S275 and S355 are presented in Table 1. These ratios were calculated using measured values presented in Wald et al. [23] and the results of 50 tensile tests documented by Byfield et al. [28] for both steel grades and the strain range up to 4%. The failure strain ratios were estimated based on the coupon tests presented in [29]. The results for the structural carbon steel ASTM A36 with the yield stress fy = 331 MPa [27] is also included in Table 1 for comparison. For the M20 grade 8.8 bolts, a simple elasticplastic model, with linear strain hardening, was applied with the characteristics presented in Table 2 and estimated based on the experimental tension tests described in [30] and the values presented in [23]. The magnitudes presented in Tables 1 and 2 allow the calculation of pairs of stressplastic strain, corresponding to the characteristic points in the diagram shown in Fig. 4 and serving as a direct input to the FE models. The strain rate effects were neglected in this study. 4.4. Joints All beams and columns including flush and fin plates are modeled using (first-order) shell elements. Such approach increases substantially a number of finite elements in the model but allows for capturing effects initiated by large structural deformations such as an inelastic bending of end plates or local buckling of compressed flanges. The bolts are represented by 1D beam elements. Additionally, there is a global single surface contact defined among

Fig. 2. Complete FE model of the entire structure. Table 2 Material properties for bolts used in the FE models. Property Modulus of elasticity Yield stress Ultimate stress Failure (total) strain Strain hardening modulus Symbol (Units) E (GPa) fy (MPa) fu (MPa) f () ESH (MPa) Magnitude 209 [30] 637 [30] 869 [23] 0.03 [31] 0.06 [30] 4074

stopping all falling elements. Depending on the abnormal loading scenario (location of the removed column), different portions of the whole building are considered, with the displacement constraints applied at the boundaries (constrained displacement in the normal direction to the vertical cutting plane). Usually, a part of the building extending two bays in each direction from the removed column is taken into account. The vertical loads recommended by GSA [3] for dynamic analysis purposes are applied in all analyses as the reference loading, LGSA = DL + 0.25LL
2

(1)

where DL = 3.65 kN/m is the design dead load, and LL = 3.5 kN/m2 is the design live load [19]. The dead load is applied in the FE models to all nodes as body force loads defined by the gravity acceleration and prescribed densities. The gravity acceleration applied to the composite slabs is increased proportionally by = 1.51 to obtain the design value DL = 3.65 kN/m2 . The portion of the live load 0.25LL = 0.88 kN/m2 is applied to the slab (only) as the additional gravity within the first second of the simulation and kept constant when the column is removed and after (see Fig. 3). Optionally, the additional gravity applied to the slabs is continuously increased beginning four seconds after the column removal to cause further damage and estimate the safety margins. Fig. 1 illustrates three considered cases of abnormal loading represented by notional removal of a column recommended

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

1227

1 0 2 5 6 9 Time [s]
Fig. 3. Time dependence of vertical loads applied to the slabs.

10

13

15

Fig. 4. Simplified stressstrain curve proposed by Galambos [27].

all the steel components represented by shell elements, e.g., flush end plates and columns, and part of the shell elements representing slabs (i.e. for thickness 130 mm). The empirical studies of beam-to-column connections indicate that the failure in the actual connections usually is initiated by the rupture of fillet welds or bolt failure (e.g., shear stripping of the threads) [31]. In the FE models, the material failure leading to the component disintegration is represented by the deletion (erosion) of a finite element from further calculations. The failure strain (especially for bolts) is another parameter strongly affecting the connections overall response and a source of the results uncertainties. 4.5. Composite concretesteel slabs Composite components, such as the reinforced concrete slab cast on steel decks considered here, are an additional challenge for numerical modeling. One difficulty is to efficiently merge responses of two materials into one representation. For small structural components, the most accurate approach is to model each material separately; for example, a concrete core with solid elements, a metal deck with shell elements, and reinforcement bars with truss or beam elements. For large-scale models, such as the ones considered herein, this strategy is infeasible today as it would result in a very large number of finite elements. The composite slab profile for the considered building and its numerical representation are presented in Fig. 5. The orthotropic slab, shown in Fig. 5, is modeled using four node shell elements and is divided into two types of strips, with different overall crosssectional properties, and positioned alternatively, side by side, along the transverse (SouthNorth, see Fig. 3) direction. The first strip has a total thickness of 130 mm and the second 70 mm. The widths for each strip are taken in a manner that conserves the original cross-sectional area (Fig. 5). Each strip is modeled as

a multilayer composite using an isotropic elasticplastic material model with different responses for tension and compression (MAT 124 [25]). Additionally, user defined through thickness integration schemes are applied, incorporating different material parameters for the layers corresponding to the concrete, steel deck, and reinforcement. All layers are represented by the same material model but have different properties for steel and concrete. The parts are divided through the thickness symmetrically about the corresponding midplanes into 16 (130 mm) and 12 (70 mm) layers, as shown in Fig. 5. For the same slab, Gillie [32] has chosen the reference plane 70 mm above the bottom of the deck, which coincides with the slab geometric centroid in the direction of the ribs. Every layer is defined by a set of properties consisting of: a local relative coordinate of the integration point zi (in the range 1 to +1), and weighting factor hi : zi = hi = Zi 0.5H Hi H

(2) (3)

where Zi is the vertical coordinate of the i-th integration point calculated as the distance from the midplane, positive in the direction of the elements normal, H is the total thickness, and Hi is the thickness of the i-th layer. An integration point is located in the middle of the layer. To conserve not only the area but also the total moment of inertia, the coordinates Zi are modified to the magnitudes Zi Hi Zi2 + Hi3 12

= Hi Zi2 ,

(4)

which, after substitution of (2) and (3), gives the formula for the modified relative coordinate zi

zi =

Zi
Hi

zi2 +

h2 i 12

(5)

All shell elements representing the slabs have normal vectors directed downward. 4.6. Material model for concrete An isotropic elasticplastic material with unique yield stress versus plastic strain curves, defined for compression and tension (MAT 124 [25]) and applicable for shell elements, was used for all layers (concrete and steel) defining the slab strips. For the concrete layer, the expected behavior is elasticplastic with failure and with varied tension and compression strengths. The basic

1228

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

Fig. 5. Composite slab profile and its numerical idealization. Table 3 Input data for concrete. Property Symbols Units Magnitudes applied in the FE model Concrete Density Modulus of elasticity Poisson ratio Compressive strength (yield stress) Tensile strength (yield stress) Maximum (failure) plastic strain in tension and compression leading to element erosion
a

(t/m3 )
E (GPa)

fc (y ) (MPa) ft (y ) (MPa)

t = c

2.00 [24] 32.5 [24] 0.2 [24] 47 [32] 3.92a 0.0035 [33]

Generated using K & C material model type 72R3.

input data for concrete layers is listed in Table 3. For this material model the same limit magnitude of the effective plastic strain for tension and compression must be used as the failure criterion. Once the failure occurs in a finite element it is deleted (eroded) from further calculations. In this way the elements experiencing extensive deformations due to material damage are removed and the structural disintegration is modeled, avoiding calculation problems [25]. The steel layers are modeled using the same material model with the material properties shown in Fig. 4 and Table 1 for steel S275, except the failure strain that was reduced here as a result of the verification procedure presented in the next section. 5. Verification and validation The following subsections describe a plan for the Verification and Validation (V & V) for finite models of the multistory

building under consideration. The whole procedure is based on the concept of hierarchical validation [34,35]. Due to the lack of experimental data (oriented toward failure at ambient temperature), the FE models developed here can be only partially validated. Herein, verification is conducted by a comparison between numerical results obtained using different modeling techniques, and validation is performed mostly as a comparison of the numerical results with published experimental data from existing relevant tests. Following the hierarchical approach, the V & V practices are integrated in four levels (tiers [35]) representing different degrees of complexity, as shown in Fig. 6. The tests classified to different levels can represent different portions of the entire structure but, more importantly, can capture different portions of the physics. The lowest level (which should be performed first) is selected as the material characterization with laboratory tests on material samples and single element numerical tests. Such tests, which are especially important for concrete, are used to approximate material properties for the whole range up to the failure and allow for estimation of the main source of uncertainties (variation of some actual material parameters can be found in [36]). At the highest level, the entire structure is analyzed but only in the elastic range. The diagram shown in Fig. 6 can be treated as a plan for hierarchical validation in which ideally the reference material should be replaced by experimental data obtained from specially designed tests on the actual structure and its components. 5.1. Material characterization and single element tests for concrete As a porous and brittle material, concrete shows a nonlinear compaction response, dramatically different strengths in tension and compression, and shear strength increased by mean

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

1229

Fig. 6. Plan for hierarchical verification and validation.

Fig. 7. Stressstrain curves for one-element compression and tension tests using materials 72R3 and 124 [25]: for compression (a), for tension (b).

stress [37]. The sophisticated material model with three invariant formulations for the failure surfaces, the so-called Karagozian & Case (K & C) Concrete Model (MAT 72R3 [37]), was applied in this study for two purposes, as input data and for verification. First, its internal automatic data generation was used to create input data for elasticplastic material model (MAT 124, see Fig. 7) applied for concrete layers in the shell elements representing reinforced slabs. Secondly, material MAT 72R3 was applied directly in detailed models of subsystems (see Section 5.4) with slabs built of solid elements, which were considered as substitutes of experiments in the third level of validation. The material 72R3 (and all its previous versions) decouples the volumetric and deviatoric responses. An equation of state defines pressure p = 1 ii as a function of volumetric strain. 3 The deviatoric stresses remain elastic until the stress point reaches the initial yield surface sy , can further increase until the maximum yield surface sm is reached and then softens to the residual yield surface sr . For increasing (deviatoric) strains beyond the initial yield surface, the stress point is located on a current failure surface s which travels between three fixed surfaces sy , sm , sr which can be

defined using eight parameters as functions of pressure p [38] sy = a0y + s m = a0 + sr = p a1f + a2f p p a1y + a2y p p a1 + a2 p

(6) (7) (8)

The parameters ai are determined by a regression fit to the available laboratory data [37]. The current surface is interpolated between the maximum failure surface sm and either the yield smin = sy or the residual failure surface smin = sr as [38] s = ()(sm smin ) + smin (9) where is a function of a modified effective plastic strain measure . Neglecting rate effects of the modified effective plastic strain , can be defined as [38]

=
0

d p

(1 + p/ft )b1

(10)

1230

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

where ft is tensile strength, and the effective plastic strain increment is given by d p = 2 3
p p ij ij .

(11)

The modified effective plastic strain (10) is used for calculation of scaled damage measure [25]

2 , ( + m )

(12)

where m is corresponding to the maximum of the function and indicates reaching the maximum failure surface sm . The contours of the scaled damage measure depict a potential crack pattern in the composite (see Fig. 13). The simplified implementation of the material 72R3 for input generation for shell elements is shown in Fig. 7. In the first step, the stress vs. (total) strain curves were generated for concrete (curve marked MAT 72R3) for the same compression strength fc = 47 MPa [32] using unconfined compression and tension virtual tests on one solid element. In the next step, the obtained relationships were simplified to the piecewise linear forms (discrete points in Fig. 7), recalculated to the form stress vs. plastic strain, and introduced as an input to the elasticplastic material [25]. As an example of verification, Fig. 7 also shows the stressstrain curves calculated for one shell element and elasticplastic material (the curve marked MAT 124). Also for comparison, Fig. 7 shows a stressstrain curve according to Eurocode 2 [33] (marked EC2) with linear extrapolation to zero stress (dashed line). Finally, the curve for compression generated by one element tested with material 72R3 was selected for further calculations in order to apply verification using detailed component models built of solid elements (see Section 5.4). For tension, two extreme magnitudes of failure strain f = 0.5 103 and f = 2.0 103 were considered for the detailed FE models used for verification in the following sections (see benchmark problems in [39] for comparison). Tension stiffening defined as gradual reduction of tensile stresses due to cracking was found to be one of the critical material parameters affecting the overall response of the composite slabs, approximated through computer simulations. It should be mentioned here that some of the parameters defining softening, in even as complex material models as 72R3, are governed by characteristic length, and are mesh-dependent [38,40]. The elasticplastic material model applied for global analyses was verified here only for two points on a failure surface; i.e. for uniaxial tension and compression. However, because in the considered global analyses the bending modes are dominant for slabs it is expected that such performance should be sufficient. This expectation is verified on the higher levels of the V & V process (see Sections 5.4 and 5.5). In contrast to accuracy, which, according to the single element tests [37] seems to be sufficient for most dedicated material models, the main computational difficulty emerging here is the stability of the calculation, especially when the analysis is supposed to reach beyond the local material damage. 5.2. Partial validation of FE models of joints Published results of tests on similar thin flush end-plate steel connections are applied here. Figs. 8 and 9 present the results of computer simulations for a major axis connection tested experimentally by Aribert et al. [31]. A HEA 360 (S355) beam is bolted on the flange of a column HEA 240 (S355) through a partial end plate (12 mm) with four M20 grade 8.8 bolts. This is a type of joint particularly convenient to use for a fast structure erection and, although designed to transmit no moments, is actually able

Fig. 8. Calculated contours of effective Mises stress for flush end-plate connection tested in [31].

200 180 160 140 120 100 80 60 40 20 0 0 20 40 60 80 100 120 140 160

Fig. 9. Comparison of experimental [31] and numerical momentrotation curves.

to carry partial bending. Fig. 8 shows the contours of effective Mises stress just before the failure, calculated for the FE model with mesh density comparable to the regular mesh applied in the current study for the global FE models. The diagram shown in Fig. 9 presents a comparison of calculated momentrotation curves with the experimental data provided in [31]. The actual tested connection was damaged due to a shear rupture of the nut threads of two first bolts while, in the FE model, failure is initiated by erosion of two beam elements for which the effective plastic strain reaches the assumed bolts failure strain f (see Table 2). The comparison of momentrotation curves shows that the FE model underestimates the initial stiffness of the connection. The initial stiffness is mostly affected by bending of the partialdepth end plate and the realization of contact between the flush plate and the column flanges. Both curves show a characteristic increase in the moment resistance when contact between the compression beam flange and the column flange occurs and the center of rotation moves to a new location. The identified modeling parameters, affecting the momentrotation relationship, are mesh density, especially for the flush end plate, failure strain f for bolts, and the contact algorithm. It should be pointed out that, although the tested beam-to-column connections are usually subject to planar bending, during a progressive collapse event more complex loading configurations with biaxial bending and torsion are present. Also, as progressive collapse can cause local reverse loading, the correct representation of unloading in the constitutive relationships defining material models is also important.

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

1231

Fig. 10. Model 2D of the slab strip cut off along the ribs. Contours of plastic strain.

5.3. Verification of FE models of slabs The 2D models are verified through comparison of ultimate bending moments with other numerical estimates calculated in [41] using the program FEAST. Numerical testing was performed for four cases on simply supported 0.6 m by 3.0 m plates of the composite slab cut off along the ribs (Fig. 10) and in the transverse direction. The plates are analyzed dynamically under four-point bending with sagging (positive) and hogging (negative) moments, caused by prescribed, time-dependent displacements applied at one third of the span. Fig. 10 shows contours of effective plastic strain for the regular 2D model of the slab with the same properties (e.g., mesh density and through the thickness integration) as in the global models. Fig. 11 presents a comparison of the response curves in terms of the momentdeflection relationships for the plates, longitudinal (along the ribs) and cut in the transverse direction, and subject to sagging and hogging moments. The shell model #2 differs with the amount of damping and loading ratio and is shown here to indicate a parametric study conducted to determine optimal modeling parameters. The maximum bending moments are in good correlation with the estimates provided in [41]. In summary, it can be stated that the assumed location of the reference plane and material behavior for concrete (Section 4.5) lead to a rather good approximation of the ultimate bending moment, especially in the primary direction along the ribs. However, the ultimate curvatures of the slab, which depend mostly on the tension softening of concrete, should be considered the main uncertainty of the global FE model (with slabs built of shell elements), requiring experimental calibration. 5.4. Subsystem level Composite floor structure Figs. 12 and 13 present a comparison of the numerical results for a selected subsystem of the structure. The subsystem contains one floor slab extended over one corner bay and around half of the surrounding bays, supported on original beams connected to the short segments of the columns. All column segments, except those of the corner column, which is unrestrained, are fixed at their bases and constrained in horizontal directions at the top. On two sides, where originally the subsystem is connected to the rest of the floor structure, the slab and supporting beams have normal displacements constrained. The homogeneous framed character of the entire structure and the assumed loading distribution determine the repeatable failure mechanisms on each floor. The chosen subsystem represents a portion of the entire structure considered the most representative part where most of the failure mechanisms develop for the corner column removal. The loading consists of regular constant gravity subject to the steel framework and increasing in time mass loads applied to the slab. Fig. 12 presents numerical results for four FE models: two regular and two detailed. As previously, the regular models have FE mesh analogues to the models used for the global analyses presented in the next section. For both regular FE models, the slab is represented by multilayer shell elements but with a different number of layers and location of the reference plane. The slab representation for regular model #1 is depicted in Fig. 13. In regular model #2, the slab is divided into 10 and 7 layers, and the reference plane is taken 70 mm above the bottom of the deck [32]. The detailed model, providing the reference results (substitute

60 Sagging - model #2 40

Moment [kNm/m]

20

Sagging - regular model

Hogging - regular model

-20 Bending along the ribs -40 0 10 20 30 40 50 Deflection [mm]

30

20 Sagging - regular model Moment [kNm/m] 10 Hogging - regular model

-10

Bending perpendicular to the ribs

-20 0 10 20 30 Deflection [mm]


Fig. 11. Comparison of momentdeflection relationships for sagging and hogging bending along the ribs and in the perpendicular direction: bending along the ribs (a), bending perpendicular to the ribs (b).

40

50

60

1.5 Lambda [ ]

0.5

0 0 100 200 300 400 500 Displacement [mm]


Fig. 12. Comparison of load paths for detailed and regular FE models of the selected representative subsystem.

of experimental data), has a higher density of FE meshes for the beams and columns and the slab modeled with solid elements and material MAT 72R3 for concrete with two magnitudes of failure strain for tension f = 0.5 103 and f = 2.0 103 .

1232

L. Kwasniewski / Engineering Structures 32 (2010) 12231235 Table 4 Comparison of natural frequencies and mass. Construction stage 1. Steel frame [42] 1st frequency mode 2nd frequency mode 3rd frequency mode 4th frequency mode Mass [t] Number of elements 0.98a EW1 1.22a NS1 1.71a 1 3.30a EW2 325 NA FE 0.994 EW1 1.259 EW2 1.547 NS1 1.548 1 368 815,124 3. Frame plus composite floors [42] 0.69 EW1 0.83 NS1 0.89 1 2.10 EW2 2302 NA FE 0.694 EW1 1.271 NS1 1.357 1 2.500 EW2 1808 1076,444

EW = EastWest direction, NS = NorthWest direction, = Rotation, 2 s ordered mode. a Experimentally tested framework with four lower steel decks.

entire building for two construction stages: the steel framework and the framework with the composite slabs. The natural (mass weighted) damping was estimated as Ds = 2 = 0.200 rad/s, (14)

Fig. 13. Concentrations of effective plastic strain and scaled damage measure in the composite slab depicting crack patterns. The state corresponds to the maximum vertical displacement of 122 mm. Regular FE mesh (a), detailed FE mesh (b).

Fig. 12 compares the load paths defined as relationships between the loading factor and the vertical displacement of the corner (unrestrained) column marked in Fig. 13. The dimensionless loading factor is defined as

where = 2.3% is the damping coefficient estimated experimentally in [43] for the first natural frequency (see Table 4) for the complete building with loading (construction stage 5), and = 2 f = 4.361 rad/s is the lowest circular frequency determined in the FE model for stage 3 (see Table 4). Global n damping results with the nodal forces Fdamp , proportional to the nodal masses mn and velocities vn [25]
damp Fn = D s m n vn .

(15)

LT LGSA

(13)

where: LT is the total vertical loading (the resultant of vertical reactions), LGSA is the reference load defined by (1) in and recommended for dynamic progressive collapse analysis by GSA [3]. Fig. 12 shows that, for the detailed model with f = 0.5 103 , the structural collapse, indicated by fast growth of the displacement, initiates under the GSA loading (1). This FE model can be considered the bottom approximation, underestimating the robustness of the subsystem. The second detailed model gives the upper approximation with the maximum loading factor reaching 1.9. Both regular models, #1 and #2, approximate ultimate loading above that required for dynamic analysis but smaller than = 2 required by GSA for static analysis [3]. Fig. 13 shows a comparison of crack patterns for regular and detailed FE models at the moment preceding the failure. The crack patterns are depicted in the slab as concentrations of plastic strain for the regular model (MAT 124) and of scaled damage measure (12) for the detailed model (MAT 72R3). 5.5. Entire structure Validation of geometry and integrity

A comparison of the numerical results with data published in [42] shows a large difference in the mass estimation for composite slabs. According to Ellis [42], the total mass of the concrete slabs in the actual building is 1997 tons while the same mass calculated based on the design drawings for eight 21 m 45 m floors without any openings is 1659 tons, which is in good correlation with the value of 1432 tons for the FE model where all openings are taken into account. The differences in mass are supposedly due to a large variation in the floor thicknesses in the actual building [36]. Although there is a good correlation between the first frequencies and modes for both stages, there are clear differences between further natural frequencies, especially the modes for the first construction stage. The reason is that the experiments were done on the framework with four lower steel decks erected [42]. In the FE model, the composite slabs change not only the frequencies but also the corresponding shapes (modes), as shown in Fig. 14. Despite some differences in the reference to the experimental data, the calculated frequencies are within the expected range. This fact indicates that all structural connections, and especially the connections between the beam ends (flush and fin plates) and the columns, are modeled properly and that the FE model shows realistic behavior in the elastics range. 6. Global analysis

To verify potential discrepancies in modeling geometry in the sense of mass and stiffness distribution, and structural integrity (connections between the components), the total mass and natural frequencies were compared with the published evaluated (for mass) and experimental (for frequencies) data [42] in Table 4. The frequencies for the global FE model were calculated using standard implicit eigenvalue analysis. The gravity loads and contacts between joined elements were initiated before the calculation of natural frequencies and corresponding modes. In the calculation scenario, in the first phase, dynamic (explicit) analysis was performed followed by intermittent (implicit) eigenvalue extraction. The calculation was carried out for the FE model of the

Depending on the location of the removed column, large portions of the entire structure are considered, covering one third or two thirds of the vertical projection. Three cases of notional column removal defined based on GSA [3] for steel frame buildings were investigated:

case # 1 loss of a column located at the corner of the building


(A1 in Fig. 1),

case # 2 loss of a column located near the middle of the long


side of the building (D1 in Fig. 1),

case # 3 loss of one interior column (E2 in Fig. 1).

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

1233

Fig. 14. Calculated first eigenmodes for the framework (top) and complete model of the framework with slabs (bottom): without slabs mode #1 (a), without slabs mode #3 (b), with slabs mode #1 (c), with slabs mode #3 (d).

The other case recommended by GSA, with loss of a column located near the middle of the short side of the building, was ignored as not appropriate for the considered structure with stiff staircases at either end of the building. All removed segments of columns are located in the first (ground) story and were selected as the most vulnerable components for the specified locations. Transient dynamic analyses were performed with the maximum simulated time period limited to 15 s. Artificially high damping was applied to reduce within such a short time period the vibrations induced by the loading and the column removal. For each case, the same loading scenario presented in Fig. 3 was applied. The structures response for case #1 is depicted in Fig. 15. Fig. 15 shows on one diagram time histories of the dimensionless loading factor defined by (13), maximum vertical displacement dz in the column section located directly above the removed segment, and the internal and kinetic energies. The numerically obtained time histories correspond well to the applied loading scenario. At t = 2 s (compare Fig. 3), the corner column is removed, and the loading is kept constant up to t = 6 s, when the loading starts to grow again. For the case #1 the structure collapses at around t 9 s, when the total loading reaches around 1.5 of the prescribed loading defined by (1). The time instance and the ultimate loading should be treated here as an approximation dependent on the loading ratio in time. To verify the estimated loading capacity, new simulations should be performed with the prescribed loading applied first and then the removed column. It is visible in Fig. 15 that the structures response under increasing loading for t > 6 s is nonlinear and most of the deformation originates in the plastic hinges and zones. For all cases, the damage did not reach the allowable extents of collapse [3] for the vertical loading defined by formula (1) which means that the analyzed building has a low potential for progressive collapse. Fig. 16 shows the load paths registered for all three cases. The abscissa represents vertical displacement dz and the ordinate the loading factor defined by (13). The maximum vertical displacement in the column after removal of its ground segment was calculated as 186, 179, and 140 mm for cases #1, #2, and #3 respectively. The failure mechanisms for three

1.5

0.5

0 0 2 4 Time [s]
Fig. 15. Time histories of loading factor , maximum vertical displacement dz in m, internal and kinetic energies in (MJ), case #1.

10

1.4 1.2 Lambda [ ] 1 0.8 0.6 0.4 0.2 0 0 100 200 300 400 500 600 Z-displacement [mm]
Fig. 16. Load factor vs. vertical displacement in the section above the removed column for the three considered cases.

considered cases are depicted in Figs. 1719 showing deformation and contours of effective plastic strain in the slabs indicating the

1234

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

Fig. 17. Deformation and contours of effective plastic strain in the slabs for case #1 at: t = 3 s (a), t = 9 s (b), and t = 10 s (c).

Fig. 18. Deformation and contours of effective plastic strain for case #2 at: t = 3 s (a), t = 8.4 s (b), and t = 9 s (c).

Fig. 19. Deformation and contours of effective plastic strain for case #3 at: t = 3 s (a), t = 8.8 s (b), and t = 9.6 s (c).

crack pattern. Three states are indicated for each case: the first one after the notional column removal, the second state just before the structural collapse propagates and the vertical loading reaches the ultimate values specified in Fig. 16, and finally the state showing structural collapse. The applied two-step loading scenario allows estimation of the safety margins defined as the ratio between the ultimate loading and the loading prescribed by formula (1). It has been assumed that the three removed columns, selected following the GSA guidelines [3], are the most vulnerable for the considered structure. However, without calculation it is difficult to predict which case gives the highest potential for progressive collapse. Also, it is not obvious if other columns (e.g. column B1, see Fig. 3) can give lower safety margin. Presented results show that the most dangerous is removal of the ground segment of the corner column which sustains less loading but is also less restrained. 7. Summary and conclusions The paper presents an attempt at building a detailed and at the same time global model for an entire large-scale structure. Such an approach can lead to computational problems due to the large number of elements, large number of computational cy-

cles required, and nonlinear structural response. In the presented work, the problem is solved through nonlinear dynamic simulations using the commercial program LS-DYNA with explicit time integration. The applied solution (explicit) method allows for taking advantage of parallel processing on multiprocessor computers and makes the computation feasible. However, this approach still requires large computational resources; for example, for the biggest FE model, the parallel calculation on 60 processors took 19 days (without mass scaling). The predictive capability of computer simulations on global FE models is evaluated using a hierarchical verification and validation program with four levels of complexity. Due to the lack of experimental data, some recommended validation tests are replaced by computer simulations using very detailed FE models of selected structural components. Of course, the credibility of such verification is much smaller than validation using specially designed experimental tests. The verification and validation program revealed two critical modeling aspects. The crucial parts in the global FE model are representations of the beam to columns connections and composite slabs. For the connections, the important modeling parameters are mesh density, especially

L. Kwasniewski / Engineering Structures 32 (2010) 12231235

1235

for fin and flush plates, and failure criteria for bolts and structural steel. A large-scale global model requires simplified representation of the slab using 2D shell elements. To encounter the variation of the cross-section and composite material, the floor slabs are divided into alternatively positioned side by side strips with two thicknesses, and modeled as a multilayer composite defined through the thickness integration. The ultimate curvatures of the slab, which depend mostly on the tension softening of concrete, should be considered the main uncertainty of the global FE model (with slabs built of shell elements), requiring experimental calibration. Global analysis was performed for one set of input data deterministically established mostly based on the literature review. For all three considered cases, the results of the global analysis were positive, showing low potential for progressive collapse of the structure. The calculation should be repeated for varied parameters to encounter the probabilistic nature of the input data such as, for example, failure strain for bolts and for concrete slabs under tension. However, such a more probabilistic approach should be more firmly based on the experimental validation. The material presented in the paper proves that with the aid of experimental validation detailed modeling can provide a realistic approximation of structural collapse. The applied modeling approach is research oriented nowadays; however, it is believed [3] that continuous software development and advancements in computer hardware will make such analysis techniques feasible for practical design in the near future. Acknowledgements Part of the presented work was done during the authors stay at the Crashworthiness and Impact Analysis Center (CIAL) at the FAMU-FSU College of Engineering. The calculation for global analyses was performed with the aid of the High-Performance Computing Centre at Florida State University. The generous support received from the CIAL and its director, Prof. Jerry Wekezer, is gratefully acknowledged. References
[1] EN 1991-1-7 (2006), Eurocode 1: Actions on structures Part 1-7: General actions accidental actions. European Committee for Standardization; 2005. [2] US Department of Defense (DoD). Unified facilities criteria (UFC), DoD minimum antiterrorism standards for buildings. Washington (DC), Department of Defense, UFC 4-010-01, US Army Corps of Engineering; 2002. p. 31. [3] General Services Administration (GSA). Progressive collapse analysis and design guidelines for new federal office buildings and major modernization projects. Washington (DC) Office of Chief Architect; 2000. [4] Marjanishvili SM. Progressive analysis procedure for progressive collapse. J Perform Constr Facilities ASCE 2004;18(2):7985. [5] Herrle KW, McKay AE. Development and application of progressive collapse design criteria for the federal government. GSA. Applied Research Associates, Inc.; 2001. http://www.buildingsecurity.us/odoc/usbuilding/pdfs/ DevAppPCDesignCriteria.pdf. [6] Bae SW, LaBoube RA, Belarbi A, Ayoub A. Progressive collapse of cold-formed steel framed structures. Thin-Walled Struct 2008;46:70619. [7] Khandelwala K, El-Tawila S, Sadekb F. Progressive collapse analysis of seismically designed steel braced frames. J Constr Steel Res 2009;65:699708. [8] Izzuddin BA, Vlassis AG, Elghazouli AY. Nethercot DA. Progressive collapse of multi-storey buildings due to sudden column loss Part I: Simplified assessment framework. Eng Struct 2008;30:130818. [9] Vlassis AG, Izzuddin BA, Elghazouli AY, Nethercot DA. Progressive collapse of multi-storey buildings due to sudden column lossPart II: Application. Eng Struct 2008;30:142438. [10] Hartmann D, Breidt M, Nguyen V, Stangenberg F, Hhler S, Schweizerhof K, et al. Structural collapse simulation under consideration of uncertainty Fundamental concept and results. Comput Struct 2008;86:206478. [11] Mller B, Liebscher M, Schweizerhof K, Mattern S, Blankenhorn G. Structural collapse simulation under consideration of uncertainty Improvement of numerical efficiency. Comput Struct 2008;86:187584. [12] Lynn KM, Isobe D. Structural collapse analysis of framed structures under impact loads using ASI-Gauss finite element method. Int J Impact Eng 2007; 34:150016. [13] Kaewkulchai G, Williamson EB. Beam element formulation and solution procedure for dynamic progressive collapse analysis. Comput Struct 2004;82: 63951.

[14] Kim HS, Kim J, An DW. Development of integrated system for progressive collapse analysis of building structures considering dynamic effects. Adv Eng Software 2009;40:18. [15] Usmani AS, Chung YC, Torero JL. How did the WTC towers collapse: A new theory. Fire Safety J 2003;38:50133. [16] Munjiza A, Bangash T, John NWM. The combined finite-discrete element method for structural failure and collapse. Eng Fract Mech 2004;71:46983. [17] Lee CH, Kim S, Han KH, Lee K. Simplified nonlinear progressive collapse analysis of welded steel moment frames. J Constr Steel Res 2009;65:1131137. [18] Mohamed OA. Assessment of progressive collapse potential in corner floor panels of reinforced concrete buildings. Eng Struct 2009;31:74957. [19] Wang YC. An analysis of the global structural behaviour of the Cardington steel-framed building during the two BRE fire tests. Eng Struct 2000;22: 40112. [20] British Steel. The Behaviour of a multi-storey steel framed building subjected to fire attack experimental data by British steel plc. 1998. http://www.mace. manchester.ac.uk/project/research/structures/strucfire/DataBase/TestData/ default1.htm. [21] BRE. Results and observations from full-scale fire test at BRE cardington. 16 January 2003. Client Report by Building Research Establishment; 2004. http://www.mace.manchester.ac.uk/project/research/structures/strucfire/ DataBase/TestData/default1.htm. [22] Bravery PN. Cardington large building test facility: Construction details for the first building. Building Research Establishment; 1994. [23] Wald F, da Silva LS, Moore D, Santiago A. Experimental behaviour of steel joints under natural fire. In: ECCS AISC workshop. 2004. [24] El-Dardiry E, Ji T. Modelling of the dynamic behaviour of profiled composite floors. Eng Struct 2006;28:56779. [25] Hallquist JO. LS-DYNA Keyword Manual. Version 971. Livermore: Livermore Software Technology Corporation; 2007. [26] ENV 1993-1-1. Eurocode 3. Design of steel structures. General rules and rules for buildings. 1994. [27] Galambos TV. Recent research and design developments in steel and composite steelconcrete structures in USA. J Constr Steel Res 2000;55(13): 289303. [28] Byfield MP, Davies JM, Dhanalakshmi M. Calculation of the strain hardening behaviour of steel structures based on mill tests. J Constr Steel Res 2005;61: 13350. [29] da Silva LS, Rebelo C, Nethercot D, Marquesa L, Simesa R. Vila Real PMM. Statistical evaluation of the lateral-torsional buckling resistance of steel Ibeams, Part 2: Variability of steel properties. J Constr Steel Res 2009;65: 83249. [30] Wheeler AT, Clarke MJ, Hancock GJ. Bending tests of bolted end plate connections in cold formed rectangular hollow sections. Research report no. R736. Australia, University of Sydney; 2006. http://www.civil.usyd.edu.au/ publications/r736.pdf. [31] Aribert J-M, Braham M, Lachal A. Testing of simple joints and their characterisation for structural analysis. J Constr Steel Res 2004;60:65981. [32] Gillie M. Development of generalized stress strain relationships for the concrete slab in shell models. PIT project research report SS2, Technical report. University of Edinburgh; 2000. http://guardian.150m.com/fire/ technicalreports.html. [33] ENV 1992-1-1:1992. Eurocode 2. Design of concrete structures Part 1: General rules and rules for buildings. 1992. [34] Kwasniewski L. On practical problems with verification and validation of computational models. Arch Civil Eng 2009;LV(3):32346. [35] Oberkampf WL, Trucano TG, Hirsch C. Verification, validation, and predictive capability in computational engineering and physics. Appl Mech Rev 2004; 57(10):34584. [36] Grantham R. Cardington static tests: Rationale and procedure. In: G.S.T. Armer, T. ODell (Eds.) LBTF, Fire, static and dynamic tests of building structures: Proceedings of the 2nd cardington conference. p. 1716. [37] Schwer LE, Malvar LJ. Simplified concrete modeling with *MAT_CONCRET_ DAMAGE_REL3. LS-DYNA Anwenderforum, Bamberg; 2005. [38] Malvar LJ, Simons D. Concrete material modeling in explicit computations. In: Workshop on recent advances in computational structural dynamics and high performance computing, USAE Waterways Experiment Station, April 2426 1996. [39] ABAQUS. ABAQUS Benchmarks Manual. ABAQUS, Inc., Version 6.6, USA; 2006. [40] Du Bois P, Feucht M, Haufe A, Kolling S. A generalized damage and failure formulation for SAMP. LS-DYNA Anwenderforum, Keynote - Vortrge II, Ulm; 2006. [41] Gillie M. Investigation of membrane flexure interaction in the Cardington slab at elevated temperatures. PIT project research report SS3, Technical report: University of Edinburg; 2000. http://guardian.150m.com/fire/ technicalreports.html. [42] Ellis BR, Ji T. Investigation of the dynamic characteristics of the steel-framed building at the Cardington. In: G.S.T. Armer, T. ODell (Eds.). LBTF, Fire, static and dynamic tests of building structures. Proceedings of the 2nd Cardington Conference. 1996. p. 18191. [43] Ellis BR. Full-scale measurements of the dynamic characteristics of buildings in the UK. J Wind Eng Ind Aerodyn 1996;59:36582.

Vous aimerez peut-être aussi