Sei sulla pagina 1di 23

Journal of Constructional Steel Research 59 (2003) 641–663

www.elsevier.com/locate/jcsr

Torsional behaviour of rectangular hollow


sections
D.J. Ridley-Ellis ∗, J.S. Owen, G. Davies
School of Civil Engineering, The University of Nottingham, Nottingham NG7-2RD, UK

Received 29 August 2001; received in revised form 22 July 2002; accepted 24 July 2002

Abstract

The results of a series of full-scale torsion tests on rectangular hollow sections (RHS) are
presented and discussed. The observed torque–twist behaviour is compared to that predicted
by an extended version of Marshall’s simplified thick wall torsion theory and finite element
(FE) analysis, and significant differences are highlighted and examined. The behaviour pre-
dicted by the FE models is shown to be identical to that predicted by Marshall’s thick wall
theory, which forms the basis of the British and European design procedures. However, even
though the experimental measurements agree with the FE and theoretical predictions in the
elastic range, the measurements of torsional capacity are significantly lower than those calcu-
lated, and this has important implications for design that may be wider than just torsion of
RHS. A number of potential causes for this behaviour are examined, but it is yet to be fully
explained. Evidence of similar behaviour in previous large-scale testing is highlighted and dis-
cussed.
 2002 Elsevier Science Ltd. All rights reserved.

Keywords: Torsion; Rectangular hollow section; Finite element; Experimental; Thick wall theory

1. Introduction
Rectangular hollow sections (RHS) are often used to resist twisting action in situ-
ations such as space frames, edge beams, grillages, and gantry stanchions because
they are inherently more efficient at resisting torsional loads than open sections. To
provide further design guidance for the use of RHS edge beams in slim floor con-


Corresponding author.
E-mail address: dan.ridley-ellis@nottingham.ac.uk (D.J. Ridley-Ellis).

0143-974X/03/$ - see front matter  2002 Elsevier Science Ltd. All rights reserved.
doi:10.1016/S0143-974X(02)00060-3
642 D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663

struction, the authors have undertaken a project that examined the behaviour of per-
forated and unperforated RHS under the actions of torsion and flexure.
The project, which compared the results of full-scale tests with the predictions of
established theory and finite element (FE) analysis, found significant anomalies for
the case of torsion of unperforated sections, despite close agreement for all other
load conditions. The results of the full-scale laboratory torsion tests suggest that RHS
have lower torsional capacities than predicted by the theory currently used as the
basis of design.
There is little published literature on the torsion testing of large-scale tubular mem-
bers, but one important previous study is that of Marshall [1], upon whose equations
the British and European design codes are based. Marshall was concerned primarily
with stiffness rather than strength, but close examination of his data has revealed
that his experimental work shows a similar deficiency in torsional resistance. These
anomalies indicate that there is an important aspect of either the structural or the
material behaviour that is not understood, which has important implications for the
design of RHS in torsion. The authors are, therefore, presenting this paper to stimu-
late debate and raise awareness of a potentially serious problem.
The paper begins with a synopsis of simplified elastic thick wall torsion theory
(Section 2) and goes on to summarise an extension of the simplified approach to
cover the elasto-plastic torque–twist relationship (Section 3). The results of a series
of full-scale torsion tests are compared to the predictions of the extended simplified
thick wall theory and FE analysis (Section 4) and it is shown that while extended
simplified thick wall theory and FE analysis are in agreement, the experimentally
measured torque–twist relationships differ significantly from the theoretical predic-
tions. A number of potential causes for this behaviour are examined (Sections 5 and
6) and evidence of similar behaviour in previous large-scale testing is highlighted
and discussed (Section 7), but it is yet to be fully explained.

2. Theoretical approximations of the elastic torsional behaviour of RHS

Theoretical approximations of the torsional behaviour of tubular sections fall into


two categories depending on how the variation of shear stress through the thickness
is treated, viz ‘thin wall’ and ‘thick wall’. Thick wall theories (e.g. Abramyan [2],
Byrne and Carré [3], and Marshall [1,4,5]) are those which attempt to account for
the variation in shear stress through the thickness and, perhaps, at the corners. Mar-
shall [4] observed that the variation in Saint Venant shear stress through the thickness
is nearly linear for typical RHS, and proposed simplified equations for the torsional
inertia constant (It, Eq. (1)) and the torsional modulus constant (Ct, Eq. (2)) that are
good approximations to his rigorous analysis. Marshall’s simplified equations are
similar in form to the classic thin wall theory of Bredt [6] in being based on the
mean perimeter (hc, Eq. (3)) and the area enclosed by it (Ah, Eq. (4)). The dimensions
of cross-section are as defined in Fig. 1.
t3hc 4tA2h
It ⫽ ⫹ (1)
3 hc
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663 643

Fig. 1. A simplified cross-section and corner detail.

hcIt
Ct ⫽ (2)
thc ⫹ 2Ah
hc ⫽ 2((b⫺t) ⫹ (h⫺t))⫺(4⫺π)(r0 ⫹ ri) (3)
(4⫺π)(r0 ⫹ ri)2
Ah ⫽ (b⫺t)(h⫺t)⫺ (4)
4
The elastic torsional capacity of an RHS Eq. (5), which is proportional to the
shear yield stress (ty), is the torque at which the external surface begins to yield and
it marks the limit of linearity in the torque–twist response. The magnitude of the
twist when the elastic torsional capacity is reached (qel) can be calculated using Eq.
(6), where L is the length of the member and G is the shear modulus of the material
Tel ⫽ tyCt (5)
LtyCt Lty hc
qel ⫽ ⫽ × (6)
GIt G 2Ah ⫹ thc

3. Theoretical approximations of the elasto-plastic torsional behaviour of


RHS

Marshall [4] was concerned primarily with elastic torsional behaviour, but his
expressions for the elastic shear stress on the internal and external surfaces can be
used to calculate the relationship between the torque and twist after yielding has
begun. This is achieved by considering the boundary of yielding as it moves inwards
through the thickness of the tube and makes use of the fact that the shear strain and
the angle of twist are directly proportional despite the yielding of the outer part of
the thickness. The yielded proportion (b) of the thickness can be calculated from
Eq. (7)
644 D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663

1 Ah tyL
b⫽ ⫹ ⫺ where 0ⱕbⱕ1 (7)
2 thc 2Gqt
The torque (T) that corresponds to this value of q can be calculated by adding
together the torque resisted by the yielded and elastic parts of the cross-section
(Eq. (8)).
T ⫽ ty(Ct,plastic ⫹ Ct,elastic) (8)
where


Ct,plastic ⫽ 2bt (b⫺bt)(h⫺bt)⫺
(4⫺π)(2r0⫺b(r0⫺ri))
4 冊 (9)

and
(1⫺b)hc,elastic((4tA2h,elastic / hc,elastic) ⫹ ((1⫺b)2t3hc,elastic / 3))
Ct,elastic ⫽ (10)
(1⫺b)thc,elastic ⫹ 2Ah,elastic
for which
hc,elastic ⫽ 2(b ⫹ h⫺2t⫺2bt)⫺(4⫺π)(r0 ⫹ ri⫺b(r0⫺ri)) (11)
and
(4⫺π)(r0 ⫹ ri⫺b(r0⫺ri))2
Ah,elastic ⫽ (b⫺t⫺bt)(h⫺t⫺bt)⫺ (12)
4
The plastic torsional capacity (Tpl), the torque at which the whole thickness of the
tube has achieved the shear yield stress (b ⫽ 1), can be estimated by considering
the flow of uniform plastic shear around the cross-section (Eq. (13)). This calculation
does not account for strain-hardened shear stresses, which may permit a higher ulti-
mate torque. The plastic torsional capacity is the same as the torsional capacity
calculated by thin wall theory.
Tpl ⫽ 2tAhty (13)
The magnitude of the twist (qpl) when the plastic torque is reached can be calcu-
lated by considering the twist required to cause shear yield at this location (Eq. (14)).
Lty hc
qpl ⫽ × (14)
G 2Ah⫺thc
Because of the assumption that the stress is uniform through the thickness, thin
wall torsion theory is not able to describe the transition from fully elastic to fully
plastic behaviour, which is a significant feature of the behaviour of RHS [7]. Thin
wall theory incorrectly predicts the range of elastic behaviour and will under-predict
the twist for a torque higher than the elastic torsion capacity. Since the thick wall
equations are no more complicated to apply than the thin wall equations, it is rec-
ommended that thick wall theory be used in design in the place of thin wall theory
for all sizes of RHS. This is already the case in the UK and Europe as the formulae
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663 645

for the torsional constants quoted in BSEN10210 [8] and BSEN10219 [9] contain
thick wall expressions that are the same as Marshall’s approximate formulae.

4. Laboratory torsion testing of RHS at Nottingham University

4.1. Introduction

This section summarises the results of torsion tests on two sizes of RHS conducted
at the University of Nottingham [10]. The tests constituted the control group for an
experimental study of the torsional behaviour of RHS with circular openings in the
web, which formed part of a wider investigation into the fundamental behaviour of
perforated RHS under torsion, bending and shear [11]. The project, which was funded
by British Steel Tubes and Pipes (now Corus Tubes) and the University of Not-
tingham, combined laboratory testing and FE modelling in order to develop the basis
for a set of design rules for perforated RHS under combined actions [12]. During
the initial stages of the project, a number of torsion tests on small-scale cold-formed
square section tube were also conducted. The results of these tests are discussed in
Section 7.

4.2. A description of the specimens, apparatus and testing procedure

The specimens tested are listed in Table 1. Four different hot-finished bars were
used in the manufacture of the specimens (two section sizes each of two different
grades). The grade S275 RHS were manufactured by British Steel Tubes and Pipes
(now Corus Tubes), and the grade S355 RHS were manufactured by Vallourec and
Mannesmann Tubes. Table 1 also includes predictions of the elastic and plastic tor-
ques (Tel and Tpl), and the corresponding rotations (qel and qpl), as calculated using
thick wall theory and measured values of the dimensions and material properties.
The torsion testing rig comprised a fixed reaction beam and a pivoted rotating
beam, between which a length of RHS could be caused to twist about its longitudinal
axis (Figs. 2 and 3). The torque was created by a matched pair of double acting
hydraulic cylinders mounted on the guide columns, and transmitted to the 2 m long
specimen through bolted connections at both ends.
The torque was measured by monitoring the shear strain on the outside surface
of a circular shaft between the specimen and the rotating beam, which ensured that
the measurements were unaffected by friction and the change in the lever arm with
rotation. The shaft was calibrated, prior to the torsion tests, by comparing the output
from the shear strain gauges with applied torques of known magnitude (up to 95
kNm).
The rotation was measured at each end of the specimen using gravity inclinometers
mounted on the connecting plates, and was, therefore, unaffected by the stiffness of
the testing apparatus and bolted connections. Three of the torsion specimens were
instrumented with electrical resistance strain gauges, and all five specimens were
painted with a brittle lime wash to increase the visibility of the Lüders lines.
646

Table 1
Specimen dimensions and thick wall theory predictions

Section Grade Measured dimensions Thick wall theory predictions

h × b × t (r0 / ri) {lweb}(mm) qel qpl Tel Tpl


(°/m) (°/m) (kNm) (kNm)

RHS 200 × 100 ×8 S275J2H 197.5 × 99.5 × 7.79(12.5/ 5.0){22} 2.14 2.73 48.1 53.7
RHS 150 × 150 × 6.3 S275J2H 149.0 × 149.0 × 6.09(11.0 / 5.0){21} 1.83 2.17 43.6 47.2
RHS 200 × 100 ×8 S355J2H 199.0 × 100.0 × 7.93(12.5 / 5.0){22} 2.26 2.90 54.6 61.1
RHS 200 × 100 ×8 S355J2H 199.0 × 100.0 × 7.93(12.5 / 5.0){22} 2.26 2.90 54.6 61.1
RHS 150 × 150 × 6.3 S355J2H 149.5 × 149.5 × 5.79(11.0 / 5.0){23} 2.20 2.57 51.8 55.7
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663 647

Fig. 2. The torsion testing apparatus (schematic).

Fig. 3. The torsion testing apparatus.

The loads applied by the hydraulic jacks were measured by load cells so that the
vertical forces could be balanced, and the movement of the end of the rotating beam
was measured by a linear potentiometer to help determine appropriate load
increments. This also provided secondary, but less reliable, measurements of torque
and rotation.
648 D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663

4.3. Finite element models

The abaqus v5.8-1 analysis program [13] was used to obtain the FE results
presented in this paper. The models of the laboratory tests were based on measured
dimensions and material properties. The latter incorporated in a non-linear material
model that allowed for the effects of strain hardening. As this investigation formed
part of a wider study of RHS with circular web openings, for which the correct
modelling of high strains around the openings was essential, four-noded shell
elements (S4R) were used. These were able to model thickness changes, while the
equivalent eight-noded shell elements (S8R) could not. However, preliminary studies
indicated that FE models of unperforated RHS in torsion gave identical results with
S4R as with S8R shells, provided a sufficient number of the former were used. The
models also included:

앫 Geometric non-linearity, as this was required for the correct modelling of the
torsional displacements;
앫 The potential for shear buckling of the tube wall;
앫 A representation of the connecting plates, and hence, a degree of warping restraint
comparable to that imposed by the testing apparatus.

4.4. A comparison of experimental results, torsion theory and FE predictions

The results of the laboratory tests were compared with the predictions of FE mod-
els and thick wall theory and the results are summarised in Table 2. The full torque
rotation relationships are shown in Figs. 4–7, for which the torque and twist are
normalised against the thick wall elastic torsional capacities. The theoretical predic-
tions are based on measured dimensions and measured values of yield stress and
Young’s modulus, but unlike the FE models they do not include an allowance for
strain hardening or shear buckling of the tube wall.
The predictions of the FE models were virtually identical to those of thick wall
theory. As is typical for hot-finished steel, the stress–strain curves for all four bars
featured a long yield plateau. This meant that the shear strains were insufficient for
significantly strain-hardened stresses to develop and that the FE predictions of tor-
sional capacity were effectively limited to the theoretical plastic torsional capacity.
None of the specimens were susceptible to shear buckling of the tube wall prior to
development of the plastic torsional capacity, although the FE model of the grade
S355J2H RHS 200 × 100 × 8 exhibited slight plastic buckling behaviour.
In contrast to the agreement between the FE and theoretical predictions, the RHS
tested in the laboratory failed to achieve even their predicted elastic torsional
capacities. The maximum torques were some 20% lower than the predicted plastic
capacities in the case of the grade S275 specimens, and some 12% lower in the case
of the S355 specimens. Despite the poor agreement with the theory, the two tests
of grade S355 RHS 200 × 100 × 8 (Fig. 6) gave similar results indicating good
experimental repeatability (less than 1% difference in capacity).
In the elastic range, the measured stiffnesses (Table 2) and shear strains (Table
Table 2
A summary of the experimental results

Section Grade Stiffness Maximum torque

Measured Percentage of Percentage of Measured Percentage of Percentage of


(kNm/°) theory FE (kNm) theory FE

RHS 200 × 100 ×8 S275J2H 12.6 112.0 111.8 43.1 80.2 79.8
RHS 150 × 150 × 6.3 S275J2H 13.4 112.7 112.4 38.8 82.2 81.8
RHS 200 × 100 ×8 S355J2H 12.3 101.6 101.6 53.2 87.0 87.0
RHS 200 × 100 ×8 S355J2H 13.1 108.2 108.2 53.6 87.7 87.7
RHS 150 × 150 × 6.3 S355J2H 11.9 101.1 99.7 49.6 89.0 88.1
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663
649
650 D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663

Fig. 4. Torque–rotation relationship for RHS 200 × 100 × 8 grade S275J2H.

Fig. 5. Torque–rotation relationship for RHS 150 × 150 × 6.3 grade S275J2H.

3) were within an acceptable tolerance (15%) of those predicted, given the difficulty
of obtaining accurate measurements of Young’s modulus with tensile tests. In the
three tests in which strains were measured, the shear strains on the external surface
were lower with respect to both the torque and twist.
In all the tests, the visible Lüders lines were almost entirely longitudinal and ran
the full length of the specimen, although a small number of transverse lines were
visible in some specimens. Examination of the specimens after the tests revealed
similar patterns on the internal surfaces. The lines, which appeared mainly on the
wall faces rather than the corners, were slightly inclined to the axis of the tube,
although there was no common pattern to the inclination or curve of the lines. These
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663 651

Fig. 6. Torque–rotation relationship for RHS 200 × 100 × 8 grade S355J2H.

Fig. 7. Torque–rotation relationship for RHS 150 × 150 × 6.3 grade S355J2H.

patterns, an example of which is shown in Fig. 8, were similar to those observed


by Marshall (Section 7).

5. Appraisal of experimental results

The experimental method, the measurements, the theoretical predictions and the
FE models were thoroughly examined in an attempt to provide an explanation for
the differences between the measured behaviour and that predicted by thick wall
652 D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663

Table 3
Experimental measurements of shear strain in the elastic range

Section Grade Torsional modulus constant, Ct Shear


strain/twist

Measured Percentage of Percentage of


(cm3) theory theory

RHS 200 × 100 × 8 S275J2H 270.8 112.1 99.9


RHS 150 × 150 × 6.3 S275J2H 259.3 113.0 99.5
RHS 200 × 100 × 8 S355J2H 269.2 108.3 92.4

Fig. 8. Typical pattern of Lüders lines.

theory and FE. The arguments supporting the validity of the experimental results
are discussed in detail elsewhere [10,11], but the main points are summarised below.

5.1. Gross measurement error

The possibility that gross measurement error was the cause of the anomalies was
thoroughly investigated by the authors, but was ruled out for the following reasons:

1. The measurements of torque, twist and shear strain in the elastic range are reason-
able, and if anything, these measurements suggest that the measured torque is
higher than the true value.
2. The measurements of torque and twist agreed with the secondary measurements
obtained from the load cells and movement of the lever arm.
3. The shear strains measured at all locations on the surface of the circular calibration
shaft were equal, and were consistent with theory for magnitudes of torque up to
95 kNm.
4. The lack of hysteresis in the measured torque–rotation relationships indicates that
the measurements were not affected by frictional losses. Furthermore, friction
between the rotating beam and the guide columns would not have affected the
torque measured on the calibration shaft.
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663 653

5. Measured torque–rotation relationships for the RHS with large web openings
agreed extremely well with the FE predictions [7,10,11].
6. The measuring and logging equipment was checked and was found to be func-
tioning within acceptable tolerances.

5.2. Buckling

The possibility that buckling of the tube walls caused the anomalies was ruled
out for the following reasons:

1. Shear buckling occurs when the shear stress in the wall reaches a critical value.
If the torsional capacity were being limited by buckling, the grade S355 specimens
would, therefore, have had similar capacities to the S275 specimens. In contrast,
the S355 specimens achieved torsional capacities that were closer to the corre-
sponding theoretical plastic torsional resistances than the S275 specimens.
2. The shapes of torque–rotation curves for sections that buckle are markedly differ-
ent from those that do not. Buckling causes a distinct peak in the curve, which
is followed by a decrease in capacity with increasing twist. In contrast, the meas-
ured torque rotation curves are of the same shape as the thick wall theory predic-
tion. Although they failed to achieve the theoretical plastic torsional resistance,
the experimental specimens were able to maintain the maximum torque with
increasing twist in all cases.
3. The FE models are capable of modelling shear buckling, yet the predicted torque–
rotation relationships show no tendency to buckle prior to development of the
theoretical plastic torsional resistance.

5.3. Boundary conditions and warping

The support conditions for an RHS in torsion can, potentially, reduce the
maximum torque it can resist. Warping restraint induces warping stresses, while a
restriction of longitudinal deformation (change in length) induces axial stresses. Both
sets of stresses act in addition to the Saint Venant shear stress with the effect of
increasing the maximum Von Mises stress in the section.
Warping restraint is unlikely to be the cause of the anomalous behaviour in the
laboratory for the following reasons:

1. According to both theory [14–17] and FE [11], square sections with constant wall
thickness do not warp, and are therefore, completely unaffected by warping
restraint. However, the square sections showed the same anomalous behaviour as
the rectangular sections.
2. According to both theory [14–17] and FE [11], the effect of warping restraint for
RHS that do warp is slight, and warping restraint has little effect on the torque–
rotation relationship.
3. A degree of warping restraint, comparable to that imposed by the testing rig, was
included in the FE models of the laboratory tests.
654 D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663

The rig was designed to allow the specimens to lengthen or shorten with twist so
as to prevent the development of axial force with increasing deformation. The FE
models of the tests also included this freedom of axial movement, but FE models
of RHS with axial restraint were found to predict identical capacities to FE models
without axial restraint. This is because the changes in length due to deformation are
very small. Furthermore, the strains measured on the surface of the specimens were
fully consistent with pure torsional shear, and the elastic response was linear.

5.4. Specimen dimensions and material properties

If incorrect dimensions or material properties had been used in the calculations,


then both the FE and theoretical predictions would have been in error and yet
remained mutually consistent. However, this is an unlikely explanation for the anom-
alies for the following reasons:

1. No single dimension reduces the theoretical elastic and plastic torsional capacities
without also affecting the elastic behaviour.
2. The specimens were re-measured after the tests and the dimensions were found
to be correct.
3. The RHS with large web openings, tested during the main part of the project,
had the same dimensions as the unperforated sections, yet the measured torque–
rotation relationships agreed extremely well with the FE predictions [7,10,11].

Similarly, there is also evidence that suggests that incorrect measurements of


material properties are not the cause of the anomalous behaviour:

1. Although Young’s modulus is difficult to measure accurately with tensile tests,


it does not influence capacity, unless there is instability.
2. Although the value of Poisson’s ratio used in the analysis (0.3) is an assumed
rather than a measured quantity, it does not influence capacity, unless there is
instability.
3. The theoretical and FE analyses were based on the assumption that there were
no residual stresses. However, hot-finished steel tubes are normalised after rolling,
and therefore, do not contain residual stresses of the same magnitude as for cold-
formed sections. Furthermore, residual stresses do not normally prevent the ability
of a section to achieve the plastic torsional capacity.

The elastic and plastic torsional capacities of an RHS are governed by the shear
yield stress of the material. In the application of thick wall theory above, and in the
FE models, the shear yield stress was calculated from the uniaxial yield stress using
the von Mises failure criterion. The uniaxial yield stress, therefore, influences the
theoretical torsional capacities without changing the stiffness, but errors in its
measurement are unlikely to be the cause of the anomalous results for the follow-
ing reasons:
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663 655

1. At least four tensile tests were conducted on material from each of the four bars,
and the measurements of yield stress did not vary sufficiently to be the sole cause
of the anomalous results (±7% for grade S275 and ±1% for grade S355).
2. Additional tensile tests of coupons inclined to the longitudinal axis of the tube,
both at Nottingham and (independently) Corus Swinden Technology Centre, did
not reveal significant anisotropy in the material behaviour (Fig. 9). The results
do suggest a slight trend of decreasing yield stress towards the transverse direc-
tion, but at less than 5% this is likely to be due to random scatter rather than a
genuine feature. Strain hardening and tensile strength were unaffected by coupon
orientation, but the yield plateau was seen to be around 25% longer in the trans-
verse direction.
3. The RHS with large web openings, tested during the main part of the project,
had the same material properties as the unperforated sections, yet the measured
torque–rotation relationships agreed extremely well with the FE predictions
[10,11].

6. Discussion of alternative shear yield criteria

For metals, the main alternative to the von Mises yield criterion, mentioned pre-
viously, is the Tresca, or principal shear stress theory, which assumes that the effect
of all the shear stresses within the material is proportional to the maximum shear
stress. The Tresca criterion relates the shear yield stress to the uniaxial yield stress
by a factor of one half, which results in a shear yield stress that is 13% lower than
that calculated by von Mises. However, experimental work in combined axial load
and torsion by investigators such as Lode [18,19], Taylor and Quinney [20], and
Hohenmser and Prager [21] have shown that the von Mises failure criterion is parti-

Fig. 9. Lack of anisotropy of uniaxial yield stress for RHS 150 × 150 × 6.3 grade S275J2H.
656 D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663

Fig. 10. The coupon for shear yield measurement.

cularly successful at modelling the behaviour of structural grade steels (for steel, the
experimentally determined ratio of axial to shear yield stress usually lies between
0.55 and 0.60, close to the von Mises value of 0.58).

6.1. Measurement of the shear yield stress

Since knowledge of the actual relationship between the shear yield stress and the
uniaxial yield stress of the hot-finished RHS was critical for this research, an attempt
was made to determine the ratio experimentally. When direct measurements are
required, the shear yield stress of materials is usually calculated from torsion tests,
a method that is clearly inappropriate in this context. Standard alternatives, such as
the Miyauchi shear test [22] and the Marciniak in-plane sheet torsion test [23],
require specialist equipment that was not available at the time of the original research.
In order to obtain measurements using the uniaxial testing equipment available, a
method was devised [11] by which the shear yield stress could be estimated from
strain measurements on a specially shaped coupon (Fig. 10).
The results are presented in Fig. 11. The shear yield stress was measured in 11

Fig. 11. Measurements of the shear yield stress.


D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663 657

separate tests on material from the same bar as the grade S355 RHS
150 × 150 × 6.3 specimen. In the figure, the measurements are normalised against
the shear yield stress as calculated from the measured uniaxial yield stress, and the
von Mises criterion. A line showing the shear yield stress according to the Tresca
criterion is also plotted. More than half (55%) of the shear coupons produced results
within 2% of the von Mises value confirming the suitability of the von Mises criterion
for calculating the shear yield stress from the tensile yield stress. Of the remaining
coupons, two (representing 18%) indicated a shear yield stresses that was signifi-
cantly higher (17 and 21%) than the von Mises value and one coupon (representing
9%) indicated a shear yield stress some 6% lower.

6.2. The role of variation in the shear yield stress

This apparent variability in the shear yield stress is a possible explanation for the
anomalous torsion test results since yielding will begin where the shear yield stress
is lowest. To study the effect of variation in shear yield stress, the FE model of the
grade S355 RHS 150 × 150 × 6.3 specimen was repeated and a randomly selected
proportion (approximately 10%) of the elements were given a reduced yield stress
(87% of the base value). Although the torque at onset of non-linearity was reduced
by 13%, the section achieved a torque within 3% of the plastic torsional capacity at
a rotation equal to the theoretical value of twist at plastic torque (Fig. 12). Despite
the fact that this and other FE models indicate that small, isolated, patches of low
yield stress do not have a strong effect on the overall torsional response, the possi-
bility that variation in the shear yield stress is the cause of the anomalous test results
cannot be firmly ruled out. This is because the shear yield behaviour in the laboratory
may be different from the simplified representation in the FE model since shear yield
in one area may promote premature shear yield in adjacent areas.

Fig. 12. The effect of random variation in shear yield stress RHS 150 × 150 × 6.3 grade S355).
658 D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663

7. Previous published experimental work

Many of the published experimental studies of the torsional properties of tubes


were conducted in the years leading to and during the Second World War on the
types of tubes found in aircraft construction. In one such series of tests, Moore and
Paul [24] measured the torsional behaviour of two tubes of rectangular section, but
found that some aspects of their behaviour were contrary to their expectations. The
work, coming a few years before Abramyan [2], relied on thin wall theory and with
the benefit of thick wall theory the results are easier to explain. The torque–rotation
curves for the two sections are shown in Fig. 13, which also shows the thick wall
theory predictions. The tubes were manufactured from an aluminium alloy that had
a yield stress in compression that was around 20% lower than that in tension. The
theoretical curves are lower bound estimates based on the compressive yield stress.
In the elastic range, the stiffness of both tubes was within 3% of the theoretical
value, but the onset of non-linearity began at around 75% of the theoretical elastic
torsional capacity. This was likely to be caused by stress concentrations at the inside
of the tight corner radii, which caused premature yielding but did not prevent the
section eventually reaching its plastic torsional capacity. In fact, due to strain harden-
ing, one of the tubes was able to resist a torque higher than the plastic capacity.
Although it is not shown in the figure, Moore and Paul indicate that buckling limits
the resistance of the more slender tube.
The results of Moore and Paul are typical of other published tests of small-scale
tubes of rectangular section in agreeing reasonably well with thick wall theory pre-
dictions. Indeed, when the authors of this paper tested cold-formed steel tubes, of a
size similar to the aluminium tubes above, the torque–rotation curves were also close
to those predicted by thick wall theory and FE models. The results of the tests, which

Fig. 13. Torque–rotation relationships obtained by Moore and Paul [24].


D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663 659

are discussed in detail elsewhere [11,25], are summarised in Fig. 14. The figure
shows the results of six separate tests and an FE prediction, for which the torque
and twist are normalised against the thick wall theory elastic torque and twist. The
measurements were in agreement with the predictions of thick wall theory and the
FE model, both in terms of elastic stiffness and plastic capacity. As is typical for
cold-formed steel, strain hardening occurred immediately after yield, and because of
this the torque resisted by the specimens continued to increase with increasing twist
until collapse occurred by plastic buckling. The FE model was able to correctly
predict this behaviour, although the buckle formed at a lower twist than in the labora-
tory.
Although previous experimental work showed a broad agreement between the
measured and theoretical behaviour of small-scale tubes of rectangular section, there
is evidence in previously published results of torsion tests of full size RHS of behav-
iour similar to that discussed in Section 4.4. Tests on hot-finished RHS, of sizes
similar to those in Section 4, were conducted by Marshall [1], who compared the
results to the predictions made by his thick wall torsion theory. Marshall measured
shear stresses on the internal and external surfaces of the RHS in addition to torque
and twist, and concluded that torsional stiffness and shear stress were adequately
predicted by his theoretical expressions. Marshall also measured the elastic and plas-
tic torsional capacities, but chose to use the Tresca yield criterion rather than the
von Mises criterion in his theoretical calculations. As a result, his theoretical predic-
tions of the torsional capacities were 13% lower than they would have been if he
had used the von Mises criterion, which is widely regarded as the more appropriate
criterion for such calculations. When Marshall’s test data were re-examined, the
results showed evidence of behaviour similar to that described in Section 4.4. Mar-
shall’s measurements of the torsional capacities are listed in Table 4, and compared
with the theoretical predictions. The calculations indicate that, although the RHS

Fig. 14. Torque–rotation relationships for small-scale torsion tests.


660

Table 4
Elastic and plastic torsional capacities for Marshall’s tests [11]

Test Measured dimensions h × b × t(mm) Grade BS4360 [26] Elastic torsional capacity Plastic torsional capacity

Measured Percentage of Measured Percentage of


(kNm) theory (kNm) theory

A 50.8 × 50.8 × 4.9 43C 2.78 91.7 3.67 101.7


B 63.5 × 63.5 × 4.9 43C 4.56 88.7 5.31 89.9
C 76.2 × 76.2 × 4.9 50 9.36 85.4 10.2 82.9
D 101.6 × 101.6 × 6.4 43C 15.9 87.6 17.7 86.6
E 101.6 × 101.6 × 4.1 43C 11.6 91.8 12.0 87.8
F 127.0 × 50.8 × 6.4 43C 8.35 82.1 10.4 86.7
G 76.2 × 38.1 × 4.1 43C 2.78 92.0 3.29 93.9
H 76.2 × 50.8 × 3.3 43C 3.54 98.3 3.75 94.0
I 101.6 × 50.8 × 3.3 43C 4.56 92.8 4.81 89.4
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663 661

behaved as predicted in the elastic range, the elastic and plastic capacities were up
to 18% lower than the theoretical predictions, and typically at least 10% lower for
the larger sections.

8. Concluding remarks

Simplified thick wall theory has been extended to describe the transition between
fully elastic and fully plastic behaviour and has been shown to be consistent with
the behaviour predicted by FE models. However, torsion tests of full size RHS indi-
cated that, although elastic behaviour was consistent with expectations, the measured
elastic and plastic torsional capacities were significantly lower than the theoretical
and FE predictions. Furthermore, a re-assessment of previously published research
has uncovered evidence of similar behaviour in the results of full-scale testing of
hot-finished RHS by the author associated with modern simplified thick wall theory
(Fig. 15). This disparity between the theoretical and measured torsional capacity is
sufficiently large to cause concern, as it indicates that current design expressions
overestimate the true capacity by around 15%. Designers need to be aware of these
limitations, but since in practice torsion does not usually govern design the conse-
quences of this shortfall in resistance are unlikely to be critical.
Although there are a number of possible explanations for this anomalous behav-
iour, no single explanation fits all the evidence available. One potential cause that
remains plausible is that a variation in the shear yield behaviour of the steel resulted
in premature yielding. However, at present there is no evidence available to support
this conjecture and its plausibility stems only through a lack of contradictory evi-
dence. The possibility that the von Mises yield criterion, which is used widely in
theoretical analysis and FE modelling is not consistent with the results of the torsion
tests has implications that are wider than just the torsional behaviour of RHS.

Fig. 15. Plastic torsional capacity—measurements vs. theory.


662 D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663

References

[1] Marshall J. Aspects of torsion of structural rectangular hollow sections. PhD thesis, University of
Strathclyde; 1972.
[2] Abramyan BL. Torsion and bending of prismatic rods of hollow rectangular section. National Advis-
ory Committee for Aeronautics, Technical Memorandum, 1319; November 1951.
[3] Byrne JG, Carré BA. Torsional stress concentrations at rounded corners of rectangular hollow sec-
tions. J Mech Engng Sci 1962;4(4):334–40.
[4] Marshall J. Derivation of torsional formulas for multiply connected thick-walled rectangular sections.
J Appl Mech 1970;37(2):399–402.
[5] Marshall J. Torsional behaviour of structural rectangular hollow sections. Struct Eng
1971;49(8):375–9.
[6] Bredt R. ‘Kritische Bemerkungen zur Drehungselastizität,’ Zeitschrift des Vereines Deutscher Inge-
nieure, 40, 1896, 785-790, 785-817.
[7] Ridley-Ellis DJ, Owen JS, Davies G. Theoretical and measured torsional behaviour of rectangular
hollow sections. In: Chung JS, Sayed M, Kashiwagi M, Setoguchi T, Hong SW, editors. Proceedings
of the 12th International Offshore and Polar Engineering Conference (ISOPE), International Society
of Offshore and Polar Engineers, vol. 4. May 2002. p. 41–7.
[8] BS EN 10210-2:1997. Hot finished structural hollow sections of non-alloy and fine grain structural
steels, part 2: tolerances, dimensions and sectional properties, British Standards Institution; 1997.
[9] BS EN 10219-2:1997. Cold formed welded structural hollow sections of non-alloy and fine grain
steels, part 2: tolerances, dimensions and sectional properties, British Standards Institution; 1997.
[10] Ridley-Ellis DJ. Owen JS, Davies G. Torsion-testing of rectangular hollow sections with web open-
ings. Report SC2000-010, The University of Nottingham; March 2000.
[11] Ridley-Ellis DJ. Rectangular hollow sections with circular web openings; fundamental behaviour in
torsion, bending and shear. PhD Thesis, The University of Nottingham; 2000.
[12] Ridley-Ellis DJ. Owen JS, Davies G. Recommendations for the design of hot-finished steel rectangu-
lar hollow sections with circular web openings; torsion, bending and shear. Report SC2000-011,
School of Civil Engineering, The University of Nottingham; July 2000.
[13] abaqus/Standard version 5.8-1, Hibbitt, Karlsson and Sorensen Inc, Pawtucket, Rhode Island,
USA; 2000.
[14] Thomas FM. An analysis of the effects due to restrained warping of thick walled box beams and
I-beams subject to torsion. PhD thesis, University of Ilinois at Urbana; 1969.
[15] Smith FA, Thomas FM, Smith JO. Torsional analysis of heavy box beams in structures. J Struct
Div 1970;96(ST3):613–35.
[16] Young WC. A practical approach to theoretical stress analysis-or where do you find it in Roark and
how do you use it effectively. Society for experimental mechanics; 1982.
[17] Megson THG. Aircraft structures for engineering students., 2nd ed. London: Edward Arnold, 1990.
[18] Lode W, Z. Physik, 36(913), 1926.
[19] Lode W. Der Einfluss der mittleren Hauptspannung auf das Fliessen der Metalle. Forsch-Arb Geb
Ing-Wes, VDI-Verlag G.M.b.H, Berlin, Heft 303; 1928.
[20] Taylor GI, Quinney H. The plastic distortion of metals. Phil Trans.Roy Soc London Ser A
1931;230:323–62.
[21] Hohenmser GI, Prager W. The mechanics of the plastic deformation of mild steel. R.T.P. Translation
number 2468, British Ministry of Aircraft Production; 1932.
[22] Miyauchi K. Stress–strain relationships in simple shear testing on in-plane deformation for various
steel sheets. In: 13th Congress of the International deep drawing research group, Melbourne, Aus-
tralia. 1984. p. 360–71.
[23] Marciniak Z. In: Aspects of metal formability. McMaster University, Ontario, Canada: Metal work-
ing research group; 1973. p. 84–91.
[24] Moore RL, Paul DA. Torsion tests of 24S-T aluminium-alloy non-circular bar and tubing. NACA
technical note 885, National Advisory Committee for Aeronautics; 1943.
[25] Ridley-Ellis DJ, Davies G, Owen JS. Fundamental behaviour of Rectangular Hollow Sections with
D.J. Ridley-Ellis et al. / Journal of Constructional Steel Research 59 (2003) 641–663 663

web openings, Proc. of the 8th International Symposium on Tubular Structures, 197-175, Singapore,
Balkema, 1998.
[26] BS 4360. Specification for weldable structural steels. British Standards Institution; 1972 [super-
seded].

Potrebbero piacerti anche