Вы находитесь на странице: 1из 8

BEHAVIOUR OF STEEL SINGLE ANGLE COMPRESSION MEMBERS AXIALLY LOADED THROUGH ONE LEG

Jie Sun and John W. Butterworth Department of Civil and Environmental Engineering, University of Auckland

SUMMARY A nonlinear finite element model applicable to steel single angle compression members eccentrically loaded through one leg has been developed using an existing finite element package. The model incorporates realistic initial geometric imperfections similar to a multi-wave local buckling mode, allows large inelastic deformations and predicts local and overall buckling behaviour. Calibration with both in-house and other experimentally acquired data showed that the model was able to predict behaviour up to ultimate load whilst using a relatively coarse mesh. A parameter study was undertaken to determine the ultimate axial load capacity of 121 equal leg struts and 88 unequal leg struts covering slenderness ratios from 30 to 300. Comparison of the results with the nominal loads prescribed by the relevant clauses of the New Zealand Steel Structures Design Standard, NZS 3404, revealed significant conservatism. A suggested interim measure for decreasing the conservatism by modifying the current interaction equation is suggested.

INTRODUCTION

Steel single angle struts are of great interest in light structures as web members and are usually connected through one leg. The resulting eccentricity due to this loading arrangement introduces end moments which are most troublesome when combined with axial compression. This complicates the buckling behaviour and creates difficulty in finding a suitable design model. The applicability of the existing design models needs to be further investigated [1, 2]. Testing of crossed diagonal angles in a 3D truss by Elgaaly et al. [3] showed that different failure modes could occur, combining local, overall and torsional effects, but that residual stress had a relatively insignificant effect on the maximum loads. A buckling mode involving buckling perpendicular to the plane of the connected leg with little twisting up to the maximum load was observed by Trahair, Usami and Galambos [4] when using fixed or hinged conditions allowing out-of-plane rotations. Bathon and Mueller [5] tested a wide range of eccentrically loaded angles using a ball joint to model end conditions unrestrained against rotation. The measured ultimate strengths were compared with the American design code. Chuenmei [6] extended the finite element analysis of eccentrically loaded angles into the nonlinear range, examining the combination of torsional-flexural buckling and local plate buckling and the interaction of overall and local buckling behaviour. Beamish and Butterworth [7] used both hybrid thin-wall beam and thin shell elements to investigate the influence of local buckling on ultimate load and post-buckling response. Both elements gave results in good agreement with each other and with experimental data.

Parameter studies using a finite element numerical model present an attractive alternative to physical testing when formulating or checking design rules for members exhibiting complex behaviour. Such studies are useful only if the numerical model is carefully checked or calibrated against results based on physical testing. The purpose of this paper is to describe the development of a nonlinear finite element model for predicting the behaviour of eccentrically loaded angles. Physical testing conducted for the purpose of calibrating the model is also described. Details are given of a parameter study aimed at establishing the ultimate axial strength of a range of eccentrically loaded angles. The ultimate loads were then compared with values derived from the relevant clauses of the New Zealand Steel Structures Standard, NZS 3404 [8]. 2 2.1 PHYSICAL TESTING Test specimens and material properties

The test struts were selected from the ordinary mild steel range supplied by BHP. The section chosen, EA 90x90x6, was based on the considerations of being a fairly typical size and having a reasonably high width/thickness ratio to encourage local buckling. Four different strut lengths were selected to cover a range of slenderness ratios from 50 to 150, generating both elastic and inelastic buckling behaviour, and to suit available test equipment. The lengths were: L=892mm (=50), L=1298mm (=73), L=1704m (=95.7) and L=2515mm (=141). NOTE - refer to last page for notation.

859

500 450 400 350 341 1 2 3 5 6 7

load increase after initial buckling up to the point of maximum load represented a modest but useful strength reserve. Maximum loads were typically in the range of 1.1-1.2 times the initial buckling loads.
Slope Strain of curve (%) Plastic strain (%) 2.02 2.14 2.34 2.74 3.34 3.94 4.74 1 2 3 4 5 6 7 11.4 5.15 3.45 2.87 2.28 2.06 1.49 2.18 2.30 2.50 2.90 3.50 4.10 4.90

Stress (Mpa)

300 250 200 150 100 50 0 0 0.16 2.18 E=200 GPa Average yield stress = 341 MPa

Strain (%)

Fig 2.1 Material model for test specimens Standard tensile tests on a number of coupons gave an average yield stress of 339MPa and an average ultimate stress of 493MPa. For comparison, the manufacturers nominal values were 260MPa and 480MPa respectively. The test data was idealised a little to give the typical stress-strain relationship shown in Fig 2.1. 2.2 Test set-up

The failure mode in seven of the test specimens involved predominant local buckling in the connected leg. This local buckling then coupled with either torsional buckling or flexural buckling about an axis parallel to the unbuckled angle leg. Most of the local buckling occurred near the end connection (tests 1, 3, 4, 5 and 7, while in tests 2 and 6 it occurred away from the connection near the mid span. Fig. 2.4 shows photographs of some failure modes.
250

209.6
200

Maximum load 182.9 159.1 153.4 145.9

179.7 172.21
150 Load (kN)

190.1

Buckling

100

50

Test 7 L=890 mm (LT mode)

Test 4 L=1298 m m (LT mode)

Test 5 L=1704 m m (LT mode)

Test 6 L=2515 mm (LG mode)

The test rig is shown in Fig 2.2. The specimens were orientated parallel to the test floor and the effect of the self-weight neglected. The ends of the specimens were bolted to pairs of back-to-back angles which were simulating the truss chord. A portion of box section strut (130x130x6) with two pairs of guides was used to apply jack loads parallel to the line joining the ends of the angle. The purpose was to prevent the loading face of the jack from rotating when the angle under test underwent large lateral deflection in the post-buckling range. Friction and play at all the interfaces was minimised by the use of close fitting greased plates and shims. The measuring system included three measurements taken at mid span of the angle specimen with one displacement transducer for the resultant vertical movement and two horizontal displacement transducers to measure lateral displacement and rotation. Another four displacement transducers were used, two at the loading face to measure the axial shortening and to check the loading face rotations, with the other two at each end of the angle to check the relative out-of-plane rotation (x). Displacement and load data was collected by a data acquisition system. In-plane rotation (y) was measured manually using a micrometer bubble level at selected points in the loading cycle. 2.3 Test Results

0 0 5 10 15 20 Axial displacement (mm) 25 30 35 40

Note: LG: Local buckling of the connected leg followed by flexural geometric axis buckling LT: Local buckling of the connected leg followed by torsional buckling

Fig 2.3 Load-axial displacement plots for EA 90x90x6 from tests 4,5,6 and 7

(a)

A total of seven struts were tested, including tests 1 and 7 with L=890mm, tests 2, 3 and 4 with L=1298mm, test 5 with L=1704mm and test 6 with L=2515mm. The experimental curves of load and axial displacement from tests 7, 4, 5 and 6 (representing the four different lengths) are shown in Fig. 2.3. It was found that the 860

(b) Fig 2.4 Typical local-overall buckling modes

Fig 2.2 Test Rig 3 3.1 NUMERICAL MODELLING Lusas nonlinear model
half-sine waves twisted about the shear centre B linear interpolation

A 3D eight-node thin shell Semiloof element (QLS8 in the Lusas [9] element library) was selected as the primary element. The general idealisation of the steel angle is shown in Fig.3.1. In order to achieve the eccentric loading through the attached leg, stiff beam elements simulating the gusset plates were connected to the shell elements at each end. The axial load was applied at the mid-point of the connected leg through these beam elements to achieve the desired eccentric compression. The nonlinear material model matched the experimentally derived stress-strain curve of Fig. 2.1 and used a Von Mise yield criterion, an associated flow rule and isotropic hardening, giving three distinct regimes - elastic, perfectly plastic and multilinear strain hardening respectively. A large displacement small strain Total Lagrangian formulation took account of the significant geometric nonlinearity. The Total Lagrangian formulation was preferred to the equivalent Updated Lagrangian formulation as it avoided the lengthy evaluations of shape function derivatives for the Semiloof elements at each load step [9,10]. Initial imperfections in the shape of several half-sine waves (resembling a typical local buckling mode) were adopted in the longitudinal direction with linear interpolation in the transverse direction as shown in Fig 3.1. The effect of residual stress was not considered in the analysis due to its insignificant effect on the maximum loads as reported by Elgaaly [3]. The solution strategies adopted for the nonlinear stepby-step response analyses involved full NewtonRaphson iteration combined with load incrementation. A restepping option was selected to accelerate the convergence. Displacement control was introduced in place of load control to avoid convergence difficulties when the solution approached a limit point [9]. A variety of other strategies including arc length control were also tried but found to be less satisfactory. 861

A (c) 3D view

shear centre (a) Viewpoint A

(b) viewpoint B first half wave Fig.3.1 Lusas Eccentric Loading Model with initial imperfection mode 3.2 Convergence studies

Mesh Density is usually an important factor influencing both the accuracy and cost of the numerical solution. Analyses to assess the effect of mesh density were performed on a typical test angle having a length of 1704mm and with both ends fixed. Initial Imperfections - In matching the mesh to the initial imperfection mode, two cases consisting of four elements per half-wave and eight elements per halfwave were considered. For the strut with 9 half waves the resulting mesh densities became 4 x 36 (coarse mesh) and 8 x 72 (fine mesh), where 4 and 8 were the transverse divisions and 36 and 72 were the longitudinal divisions. The numerical solutions and the physical test results (from Section 2) are shown in Fig 3.2. Nearly identical solutions were obtained from the coarse and fine mesh models, with the ultimate load capacities matching the test buckling loads with reasonable accuracy, having errors of 2.1% and 1.4% respectively. However, as can be seen there was considerable difference in the post-buckling range.

Load capacity (kN)

200 150 100 50 0

150.1

151.2

153.4

4x36

8x72

Test data

Apart from the apparently anomalous result for the 2515mm specimen (discussed later), numerical predictions using an initial imperfection amplitude of e=0.333t gave the best agreement with the test buckling loads, with errors in the range 2.5%.
Effect of the magnitude of the initial geometric imperfections for test angles EA 90x90x6
post-buckling strength reserve
L=1704 mm Fy=341 MPa one leg fixed wave number: 9 (direction1) with magnitude of 0.333t disp. norm: 0.001

200 180 160 140


200

Load (kN)

120 100

buckling load

2.50 Errors (%)

2.00 1.50 1.00 0.50 0.00 4x36

errors (%)

80 60 40 20 0 0 1 2

2.15 1.43

180

Load capacity (kN)

mesh size: 8x72 mesh size: 4x36 Test Results

160

8x72

140

straight column Magnitude = 0.167t Magnitude = 0.333t Magnitude = 0.5t Magnitude = 0.667t

10

4.0 3.0 2.0 1.0 0.0 -1.0 -2.0 -3.0 -4.0

2.5

L=890mm L=1298mm L=1704mm

-2.45

-2.09

120

Axial displacement (mm)


100

Magnitude = 0.333t mm

Test results

Fig 3.2 Comparison of analysis and test data Convergence criteria - tight tolerances were required to maintain control of the analysis in the presence of significant geometric nonlinearity. The Euclidean displacement norm,

80 0 500 1000 1500 Length (mm) 2000 2500 3000

Fig 3.4 Comparison of Lusas imperfect models with test data

r a d = r 2 x 100 a2

(5.1),

one of a number of convergence measures available in Lusas, was generally used. The results obtained by using criteria of d = 0.1 ~ 0.0005 are presented in Fig 3.3. Solutions for the ultimate loads varied little when using different tolerance factors. The maximum error of 2.8% compared to the test data was found at d = 0.01, however, with the smaller factors used, identical solutions occurred between d = 0.001 and d = 0.0005.
160 140 120 100 80 60 40 20 0 149.8 149.1 150.1 150.1

Direction of the initial wave - Analyses to assess the effect of the direction of the assumed imperfection wave were performed by imposing waves with opposite directions, 1 and 2, as defined in Fig 3.5. Higher capacities were obtained when using direction 2 for angles of L=890 ~ 1704mm. In the post-buckling region, significant difference was observed for the shortest angle with the difference larger than the effect on the ultimate loads, but the influence decreasing with member length. For the most slender member, identical solutions were obtained for both directions.
first half sine wave twisted about the shear centre in anticlockwise direction first half sine wave twisted about the shear centre in clockwise direction

3.00 2.50 Errors (%) 2.00 1.50 1.00 0.50 0.00 Disp. norm=0.1 2.35

2.80 2.15 2.15

Load Capacity (kN)

Disp. norm=0.1

Disp. norm=0.01

Disp. Disp. norm=0.001 norm=0.0005

Disp. norm=0.01

Disp. norm=0.001

Disp. norm=0.0005

(a) Load capacity

(b) Errors

Fig 3.3 Convergence criteria study


(L=1704mm, Pcr=153.4 kN)

(a) direction 1

(b) direction 2

Fig 3.5 assumed wave directions


Load-axial displacement plot for EA 90x90x6 (Lusas analysis)

3.3

Initial imperfection effect


200 180 160 140 Load (kN) 120 100 80 60 40 20 0 0

Amplitude of the initial imperfection wave Numerical predictions for the test specimens using initial imperfection amplitudes of 0.167t, 0.333t, 0.5t and 0.667t are summarised in Fig 3.4, together with the existing physical test results. The greatest difference in ultimate loads occurred between a straight column and the corresponding imperfect column with smaller differences resulting from the different imperfection amplitudes. The differences were largest for the shortest column (L=890mm), with the sensitivity decreasing with increasing length of angle. 862

Fy=341 MPa one leg fixed assumed wave number of 5,7,9,13 with max. magnitude of 0.167t disp. norm: 0.001 L=890mm (direction 2), Pu=188.2kN L=890mm (direction 1), Pu=184.4kN L=1298mm (direction 2), Pu=175.4kN L=1298mm (direction 1), Pu=170.9kN L=1704mm (direction 2), Pu=153.8kN L=1704mm (direction 1), Pu=150.1kN L=2515mm (direction 2), Pu=97.6kN L=2515mm (direction 1), Pu=99.84kN

Axial displacement (mm)

Fig 3.6 Effect of imperfection direction

3.4

Numerical buckling behaviour

amplified at the connected leg rather near mid span, as shown in Fig 3.11. (2) post-buckling
stage (1) prebuckling stage

Ultimate load capacity and buckling response - the effect of end fixity was evaluated by applying three end conditions to the established model Fixed end both the in-plane (y) and out-of-plane (x) rotations fixed Hinged end out-of-plane rotation released Simple end both rotations released. The results (using direction 1 imperfections in all cases) are summarised in Table 3.1. Fixed end provided load capacities of 20% ~ 42% higher than the simple end and the hinged end gave ultimate loads of 7.5% ~ 13% lower than the fixed end. The significant stiffening effects due to end fixity on both ultimate loads and post-buckling response is shown in Fig 3.7. The effect on ultimate load capacity decreased with increasing slenderness as shown in Fig 3.8.
Ultimate load (kN) fixed hinged simple end end end 184.4 159.8 130 170.9 149.5 123 150.1 133 112 99.8 92.4 83.3

(c) viewpoint C

amplified local waves

C vertical bending B

(b) viewpoint B twist effect

(d) 3D view A

(a) viewpoint A

strut length (mm) 890 1298 1704 2515

(3) ultimate stage

Fig 3.9 Buckling mode under fixed or hinged end condition

Table 3.1 Effect of end fixity


200 180 160 140 120 100 80 60 40 20 0 0 2 4 Axial displacement (mm) 6

Fig 3.10 Interaction in a slender member


(2) post-buckling stage

Load capacity (kN)

200 150 100 50 0 0 100 200 Slenderness ratio L/rv fixed end hinged end simple end

Load (kN)

fixed end hinged end simple end

(1) prebuckling stage

(c) viewpoint C Fig 3.7 Typical Load-axial displacement plot Fig 3.8 Effect of end fixity horizontal bending C B

amplified local waves

Buckling modes - The finite element results showed predominant vertical rather than horizontal deformation with the angle buckling about an axis parallel to the connected leg as shown in Fig 3.9. NOTE: the connected leg was horizontal in the finite element model, but vertical in the physical tests. Twist was almost imperceptible in the pre-buckling stage, but dominated the post-buckling region. The local waves were amplified and interacted with the overall bending, with the effect more pronounced in the stockier members, as shown in Fig 3.10. However, under simple end conditions the angles exhibited quite different modes. Horizontal bending increased rapidly and became dominant at the ultimate loading stage in the post-buckling region. The angle buckled about a major principal axis first and was prone to bend about the axis perpendicular to the connected leg at the ultimate loading stage. Twist was insignificant throughout the response and local waves were 863

(b) viewpoint B

(d) 3D view

(a) viewpoint A

(3) ultimate stage

Fig 3.11 Buckling mode, simple end condition Parameter Study - Analyses were conducted on eleven equal leg angle sections and eight unequal leg angle sections using properties taken from the BHP Specification [11]. Eleven slenderness ratios were considered for each different angle section. For the first nine, an imperfect model with imperfection wave amplitude of 0.333t was adopted. For the two highest slenderness ratios imperfections were set to zero as

their effect was known to be negligible. The load was applied at the mid point of the connected leg. The hinged boundary condition was used, allowing out of plane rotations. The results are summarised in Table 3.2. 4 COMPARISON

greater extent than the ultimate loads. The experimental specimens had a range of (undetermined) imperfections causing these behaviour variations. Another item of interest was the disparity in the experimental buckling load of the most slender angle, which exceeded the analytical value by 35.2%. A possible explanation was that the initial imperfection resulted in the angle bowing away from its preferred buckling direction causing it to follow a different equilibrium path leading to a higher limit point from which it buckled and reverted to its preferred path. This type of behaviour would also be expected to show the observed steeper fall off in the immediate post-buckling regime [12]. A similar effect can be seen in Fig. 4.1 where the imperfection directions 1 and 2 give contrasting post-buckling slopes, although relatively small differences in buckling load.

Compared with the physical test results, the analyses based on a relatively coarse mesh predicted response up to the ultimate load level reasonably closely and permitted appropriate choices for initial imperfection parameters to be made. Some disagreement was apparent in the post-buckling region and the theoretical buckling modes did not give satisfactory coincidence with experimental observations. It was thought that this disagreement was due to the assumed imperfection modes affecting the post-buckling responses to a

Slenderness ratio (L/rv) versus theoretical ultimate loads Pu (kN) L/rv - Pu L/rv - Pu
2

L/rv - Pu

L/rv - Pu
2

L/rv - Pu

L/rv - Pu
2

L/rv - Pu

L/rv - Pu
2

L/rv - Pu

L/rv - Pu
2

EA 150x150x19
(A=5360 mm , Fy=280 MPa) 30 50 70 90 110 130 -

EA 150x150x10
(A=2790mm , Fy=320 MPa)

EA 100x100x12
(A=2260 mm , Fy=260 MPa)

EA 100x100x6
(A=1170 mm , Fy=260 MPa)

EA 45x45x6
(A=506mm , Fy=260 MPa)

837 822 807 784 633 556

150 170 190 240 300 -

504 443 385 303 210

30 50 70 90 110 130 -

442 422 400 372 314 273

150 170 190 240 300 -

236 202 174 129 90

30 50 70 90 110 130 -

326 322 317 308 250 229

150 170 190 240 300 -

206 179 156 125 87

30 50 70 90 110 130 -

142 140 134 125 110 98

150 170 190 240 300 -

86 75 65 50 35

30 50 70 90 110 130 -

74 73 72 70 57 51

150 170 190 240 300 -

47 41 36 29 20

EA 90x90x10
(A=1620 mm , Fy=260 MPa) 30 50 70 90 110 130 2

EA 90x90x6
(A=1050mm , Fy=260 MPa) 30 50 70 90 110 130 2

EA 75x75x10
(A=1340mm , Fy=260 MPa) 30 50 70 90 110 130 2

EA 75x75x5
(A=672mm , Fy=260 MPa) 30 50 70 90 110 130 150 - 56 95 170 - 49 91 190 - 43 87 240 - 33 83 300 - 19 72 64 UA 125x75x12 (A2260mm , Fy=260 MPa) 30 50 70 90 110 130 2 2

EA 45x45x3
(A=263mm , Fy=260 MPa) 30 50 70 90 110 130 150 - 20 34 170 - 18 33 190 - 15 31 240 - 12 30 300 - 8.2 26 23 UA125x75x6 (A1170mm , Fy=260 MPa)
2 2

243 150 - 151 240 170 - 133 235 190 - 117 227 240 - 92 188 300 - 64 167 EA 55x55x5
(A=489mm , Fy=260 MPa)
2

134 150 - 81 131 170 - 70 125 190 - 61 118 240 - 47 102 300 - 33 92 UA 150x100x12
(A=2870mm , Fy=300 MPa)
2

206 150 - 131 204 170 - 113 201 190 - 101 196 240 - 80 159 300 - 56 142 UA 150x100x10
2

(A=2300mm , Fy=320MPa)

30 50 70 90 110 130 -

73 71 68 67 56 50

150 170 190 240 300 -

43 39 34 27 19

30 50 70 90 110 130 -

381 375 372 367 362 357

150 170 190 240 300 -

341 316 283 230 163

30 50 70 90 110 130 -

329 326 322 315 310 301

150 170 190 240 300 -

283 260 232 184 129

257 255 253 251 248 248

150 170 190 240 300 -

246 242 229 198 149

30 50 70 90 110 130 -

120 119 118 116 115 113

150 170 190 240 300 -

110 106 100 88 64

UA 100x75x10
(A=1580mm , Fy=260 MPa) 30 50 70 90 110 130 2

UA 100x75x6
(A=1020mm , Fy=260 MPa)
2

UA 75x50x8
(A=921mm , Fy=260MPa)
2

UA 75x50x6
(A=721mm , Fy=260 MPa)
2

208 205 203 201 199 194

150 170 190 240 300 -

183 167 148 120 86

30 50 70 90 110 130 -

118 116 114 112 109 105

150 170 190 240 300 -

98 90 80 63 44

30 50 70 90 110 130 -

113 112 111 110 110 109

150 170 190 240 300 -

107 101 91 78 58

30 50 70 90 110 130 -

83 82 81 81 79 79

150 170 190 240 300 -

77 73 67 56 41

Table 3.2 Parameter study - ultimate axial loads for equal and unequal angles 864

Column Strength Curve for EA 90x90x10

250
imperfect column with direction 1

450 400

straight column

Load Capacity N* (kN)

200

imperfect column with direction 2

Lusas predictions Clause 6.6 Le=0.85L Clause 6.6 Le=L Clause 8.4.6 Le=0.5L Clause 8.4.6 Le=L

350 300 250 200 150 100 50 0 0 50 100 150

Load (kN)

150

100
L=890mm L=1298mm L=1704mm L=2515mm

50

0 0 5 10 15 20

200

250

300

350

Axial displacement (mm)

Slenderness ratio (L/r v)

Fig 4.1 Effect of imperfection modes 5 RELEVANCE TO NZS 3404

Fig 5.2 Typical column strength curve (EA)


Column Strength Curve for UA 75x50x6 200

Load Capacity N* (kN)

Two design models are currently recommended in the NZ Steel Design Standard, NZS 3404 [8]. Clause 6.6 determines N c for compression alone and is only applicable to members with 150, eq (5.1).
N N c
*

180 160 140 120 100 80 60 40 20 0 0 50 100 150

Lusas predictions Clause 6.6 Le=0.85L Clause 6.6 Le=L Clause 8.4.6 Le=0.5L Clause 8.4.6 Le=L

(5.1)

200

250

300

350

In clause 8.4.6, the beam-column model is allowed for designing for a combination of moment and axial load (eq 5.2 and Fig 5.1). The Le effect is only considered in evaluating Mbu, and the pin-end (Le=L) is allowed to determine Nch.
N
*

Slenderness ratio (L/r v)

Fig 5.3 Typical column strength curve (UA)


Load ratio vs. moment ratio for Lusas prediction as per NZ 3404 provisions (clause 8.4.6, Le=0.5L) 1.2 EA 90x90x6 EA 90x90x10 EA 100x100x6 EA 100x100x12 EA 150x150x10 EA 150x150x19 EA 75x75x5 EA 75x75x10 EA 45x45x3 EA 45x45x6 EA 55x55x5 1.8

Load ratio N*/Nch

N ch

Mh

1 0.8 0.6
Modified Equ. 5.3

M bu cos

(5.2)

0.4 0.2 0
Equ. 5.2

0.2

0.4

gusset plate

Moment ratio M*/(Mbu*cos( ))

0.6

0.8

1.2

1.4

1.6

h
u

h
v

Fig 5.4 Load - moment interaction (EA)


e
Load ratio vs. moment ratio for Lusas prediction as per NZ 3404 provisions (clause 8.4.6, Le=0.5L) 1.2
UA 75x50x6

Fig 5.1 Design model in clause 8.4.6


Load ratio N*/Nch

1 0.8 0.6
Modified Equ. 5.3

UA 75x75x8 UA 100x75x6 UA 100x75x10 UA 125x75x6 UA 125x75x12

The corresponding column strength curves using = 1.0 with effective length factors of 1.0 and 0.85 for clause 6.6 and of 1.0 and 0.5 for clause 8.4.6 are displayed together with the analytical ultimate loads in Table 3.2. Typical results are plotted in Fig 5.2 for an equal leg angle (EA) and in Fig 5.3 for an unequal leg angle (UA). Clauses 6.6 underestimated capacities in the high slenderness ratio range. Similarly, clause 8.4.6 underestimated capacities in the low slenderness ratio range. Clause 8.4.6 was also relatively insensitive to effective length and consequently unable to reflect the real effect of the end conditions even though the end restraint may exert a strong influence on the load capacity for struts of low to medium slenderness.

0.4 0.2 0 0 0.2 0.4


Equ.5.2

Moment ratio M*/(Mbu*cos( ))

0.6

0.8

1.2

1.4

1.6

1.8

Fig 5.5 Load ratio- moment interaction (UA) The combination of axial loads and bending moments on the single angle struts for use with clause 8.4.6 are shown in Figs 5.4 and Fig 5.5 for equal and unequal angles respectively. Axial load ratio (N*/Nch) is shown on the Y-axis and the moment ratios (M*/Mbucos) on the X-axis. The majority of the numerical results are scattered above the interaction equation, with different

865

scatter patterns in the EA and UA groups. Significant conservatism in the interaction equation is apparent. Because of the scatter it is clear that the present interaction equation can not be simply altered to produce a good fit. However, limited improvement could be achieved by applying clause 8.4.6 over the full range of slenderness ratios, retaining the linear interaction formula in its present form and imposing different scaling factors for interaction. This was done by rotating the interaction line and moving it so that it formed a closer lower bound to the groups of calculated capacities. The resulting modified equation became:
N* 0.8 N ch
+

2.

M* 17 . M bu cos

=1

(5.3)

This gave increases in axial load capacity in the range 4.1% to 15.4% for the groups of angles used in the parameter study when compared with the old interaction equation. The range of slenderness ratios considered ranged up to =190. A more detailed improvement would require the incorporation of more parameters, such as the B/t ratio, in the calculation of Nch and Mbu, together with a nonlinear interaction equation. 6 CONCLUSIONS

A nonlinear finite element model applicable to steel single angle compression members was able to predict local and overall buckling behaviour under eccentric axial loading through one leg up to ultimate load whilst using a relatively coarse mesh. Calibration against physical tests enabled suitable waveforms and amplitudes of initial imperfections to be chosen in order to predict initial buckling and ultimate loads with good accuracy. An amplitude of 0.333t for the initial imperfection waves combined with a mode 1 form generally gave the best match to test results. Post-buckling behaviour, appeared to be particularly sensitive to both magnitude and direction of initial imperfections making it more difficult to predict. Comparison of the results of a parameter study giving the axial strengths of over 200 equal and unequal angles with the nominal axial strength prescribed by the relevant clauses of the New Zealand Steel Structures Design Standard, showed significant conservatism. The inclusion of two additional scaling factors in the current code interaction equation resulted in increases in axial capacity of the order of 10%. 7 1. REFERENCES Temple, M.C. and Sakla, S.S., Consideration for the Design of Single Angle Compression Members Attached by One Leg, Proceedings, International Conference on Structural Stability and Design, Kitipornchai, Hancock, Bradford (Eds), Balkema, Rotterdam, 1995, pp107-112.

NZ Heavy Engineering Research Association, Steel Design and Construction Bulletin, No. 17, December 1995. 3. Elgaaly, H., Dagher, and Davids, W., Behaviour of Single Angle Compression Members, Journal of Structural Engineering, Vol.117 No12, December, 1991, ASCE pp3720~3741 4. Trahair, N.S., Usami, T. and Galambos, T.V., Eccentrically Loaded Single Angle Columns, Research Report No. 11, Dept. of Civil and Environmental Engineering, Washington Univ., St Louis, Mo., Aug., 1969. 5. Bathon, L, Mueller, W.H. and Kempner, L., Ultimate Load Capacity of Single Steel Angles, Journal of Structural Engineering, Vol No1, ASCE 1993, pp 229~300 6. Chuenmei, G., Elastoplastic Buckling of Single Angle Columns, Journal of Structural Engineering, Vol. 110, No6, June, 1984, ASCE, pp 1391~1395 7. Beamish, M.J., and Butterworth, J.W., Inelastic local and lateral Buckling of Thin-Walled Steel Members, Report No 494, Dept. of Civil Engineering, University of Auckland, 1991. 8. Steel Structures Standard, NZS 3404:Part 1:1997, Standards New Zealand, Wellington. 9. Lusas Finite Element Analysis System - Theory Manual, Version 10.0, 1990 Finite Element Analysis LTD, UK 10. Javaherian, H., Dowling, P.J., Lyons, L.P.R., Nonlinear Finite Element Analysis of Shell Structures Using the Semi-loof Element, Computers and Structures, Vol 12, pp 147-159. 11. Hot Rolled and Structural Steel Products, 1994 Edition, BHP Co. Pty Ltd, Melbourne. 12. Thompson, J.M.T. and Hunt, G.W., A General Theory of Elastic Stability, Wiley, 1973.

8
A Fy Mbu N* Nch Pu t d

NOTATION
cross section area yield stress nominal moment capacity about major principal u axis design axial force nominal compression capacity (about h axis) ultimate load thickness of angle leg Euclidean displacement norm (tolerance factor) total displacement in-plane rotation about y axis in Fig 3.1

E Le Mh* Nc Pcr rv

elastic modulus effective length end moment due to eccentricity = eN* nominal compression capacity experimental buckling load radius of gyration about minor principal v-axis slenderness ratio about minor principal v-axis displacement increment strength reduction factor out-of-plane rotation about y axis in Fig 3.1

r a
y

r a
x

866

Вам также может понравиться