You are on page 1of 37

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/344371029

Experimental study on material properties and stress concentration factors of


stainless-steel hybrid tubular joints

Article  in  Construction and Building Materials · September 2020


DOI: 10.1016/j.conbuildmat.2020.121103

CITATIONS READS

5 205

6 authors, including:

Ran Feng Krishanu Roy


Harbin Institute of Technology, Shenzhen University of Auckland
121 PUBLICATIONS   1,280 CITATIONS    148 PUBLICATIONS   2,450 CITATIONS   

SEE PROFILE SEE PROFILE

Boshan Chen James B.P. Lim


Tsinghua University University of Auckland
43 PUBLICATIONS   733 CITATIONS    290 PUBLICATIONS   5,022 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Joints and angular connections View project

Stability of Structures under their Own Weight View project

All content following this page was uploaded by Boshan Chen on 24 September 2020.

The user has requested enhancement of the downloaded file.


Experimental study on material properties and stress concentration
factors of stainless-steel hybrid tubular joints
Ran Feng a, Zhipeng Huang b, Zhenming Chen a,c, Krishanu Roy d*, Boshan Chen d*,

James B.P. Lim d


a
School of Civil and Environmental Engineering, Harbin Institute of Technology, Shenzhen, China (518055)
b
China Energy Engineering Group Anhui Electric Power Design Institute CO., LTD., Hefei, China (230601)
c
China Construction Science and Industry Corporation LTD, Shenzhen, China (518054)
d
Department of Civil and Environmental Engineering, The University of Auckland, New Zealand

Abstract

This paper presents an experimental investigation to determine the stress concentration factors
(SCFs) of stainless-steel hybrid tubular joints, connecting circular hollow sections (CHSs) with
square hollow sections (SHSs) under axial loads. Austenitic stainless-steel (AISI 304 grade) was
used to manufacture all specimens. In total, twelve experiments were conducted comprising T, Y
and X joints. To obtain the material properties of stainless-steel hybrid tubular joints, tensile coupon
tests were conducted. For each joint, eight unidirectional strain gauges were placed to measure the
hot spot strain perpendicular to the weld toe and the strain parallel to the weld toe. To analyze the
strain gauge data from experiments, the hot spot stress method was used, adopting the quadratic
extrapolation method, to determine the SCFs of stainless-steel hybrid tubular joints. From the
results of experiments, strain concentration factors and conversion coefficients were recommended
for stainless-steel hybrid tubular joints. The maximum SCFs recorded from the experiments on the
SHS chord and CHS brace of stainless-steel hybrid tubular joints were then compared against the
design values calculated in accordance with the CIDECT Guideline 8 and the proposed design
formulas by Tong et al. and Yin et al. for carbon steel hybrid tubular joints. It is shown that the
current design formulas as prescribed in CIDECT Guideline 8, are conservative by around 20%,
while determining SCFs of such stainless-steel hybrid tubular joints.

Keywords: Circular hollow section (CHS); Hybrid; Square hollow section (SHS); Stainless-steel;

*Corresponding author. Tel.: +64 2108840683 (B. Chen); +64 223917991 (K. Roy).
E-mail address: bche719@aucklanduni.ac.nz (B. Chen); kroy405@aucklanduni.ac.nz (K. Roy).
1
Stress concentration factor (SCF)

1. Introduction

Structural applications of stainless-steel are increasing steadily [1-4] and the use of
stainless-steel hybrid tubular sections is becoming increasingly popular in offshore structures. The
joints of these tubular sections, when used in offshore structures, can suffer from severe corrosion
and fatigue failure under cyclic loading [5]. Attention should thus be paid to fatigue failure
mechanism of stainless-steel tubular joints, particularly at locations of local high stress. It is
recommended under fatigue design specifications for welded tubular structures that fatigue
behaviour can be evaluated by the hot spot stress method [6]. The hot spot stress is the peak
geometrical stress at the tubular joints where splits initiate. The actual idea of the hot spot stress
method is to determine the hot spot stress amplitude of the joint, and then substitute the
corresponding type of S-N curve, so as to obtain the fatigue life of the structure or specimen under a
certain stress state [6]. However, the hot spot stress is not easily measured in engineering situations
such as tubular joints [7]. The literature review also reveals that no research is available on fatigue
failure mechanism of stainless-steel hybrid tubular joints.
This paper presents an experimental study on stress concentration factors (SCFs) of
stainless-steel hybrid tubular joints connecting circular hollow sections (CHSs) and square hollow
sections (SHSs) (Fig. 1). In total, twelve laboratory tests were conducted on T, Y and X joints.
Austenitic stainless-steel (AISI 304 grade) was used for all test specimens. Strain concentration
factors and conversion coefficients are recommended from the results of experiments.
The only work available in the literature for stainless-steel tubular joints is by Feng and Young
[7], who studied the SCFs of stainless-steel SHS and RHS tubular X joints. From the results of their
study [7], they developed a unified design equation for SCFs of stainless-steel tubular X joints. In
the existing CIDECT (International Committee for the Development and Study of Tubular
Structures) Guideline 8 [8] and DNVGL-RP-C203 [9], the design formulas refer only to the SCFs
of carbon steel tubular joints and no information is given on the SCFs of stainless-steel hybrid
tubular joints. The formulae given in CIDECT Guideline 8 [8] are same as that given in
DNVGL-RP-C203 [9]. Therefore, the results calculated by CIDECT Guideline 8 [8] were used in
this paper.

2
Significant research is, however, available in the literature for SCFs of carbon steel circular,
square and rectangular hollow section (CHS, SHS and RHS) tubular joints. Toprac et al. [5]
experimentally investigated the effects of cyclic loading on the fatigue behaviour of welded T joints
and found that the key factors which affected the fatigue behaviour of such joints is the SCFs. On
the other hand, Kuang et al. [10], Gibstein [11], Gho and Gao [12], Woghiren and Brennan [13]
conducted finite element (FE) investigations on the SCFs of CHS tubular T and Y joints,
overlapped tubular K(N) joints and multi-planar tubular KK joints. Ahmadi et al. [14] carried out
both experiments and FE analysis to investigate the SCFs of KT joints. A study on the hot spot
stress distributions of tubular KK-joints under axial loads was conducted by Yue et al. [15]. The
results revealed that the stress distributions of tubular KK-joints are quite different from those of
uniplanar T and K joints. Ahmadi and Zavvar [16] investigated the SCFs of multi-planar tubular KT
joints under bending. Furthermore, the fatigue behaviour of CHS-to-SHS tubular T and Y joints was
studied by Mashiri et al. [17], Tong et al. [18] and Yin et al. [19].
In terms of carbon steel hybrid tubular joints, Tagoe [20] studied the fatigue behaviour of
hybrid tubular joints, including CHS brace-to-SHS chord (CHS-to-SHS) tubular joints and SHS
brace-to-CHS chord (SHS-to-CHS) tubular joints, which are quite different from those of CHS,
SHS and RHS tubular joints. Gandhi and Berge [21] conducted an experimental investigation on
SCFs of hybrid CHS-to-SHS tubular T joints and concluded that the SCFs of hybrid CHS-to-SHS
tubular T joints are lesser than the SCFs of SHS tubular T joints. Furthermore, Bian and Lim [22]
conducted an experimental study on hollow section T joints under compression and bending. The
experimental results were statistically evaluated, and the SCF values for CHS-to-RHS hollow
section joints were found to be in between of those for CHS-to-CHS and RHS-to-RHS hollow
section joints [22].
The material properties of carbon steel are significantly different from stainless-steel.
Compared with carbon steel, stainless-steel has a comparatively low proportional limit stress with
no distinct yield plateau and pronounced strain hardening behaviour. Therefore, the design formulas
for SCFs of carbon steel tubular joints may not be suitable for SCFs of stainless-steel tubular joints.
In addition, the current CIDECT Guideline 8 [8] does not provide the design formulas for SCFs of
hybrid tubular joints.
As mentioned previously, this paper presents a detailed experimental study on SCFs of
stainless-steel hybrid tubular joints. Twelve specimens of four T, four Y and four X joints were
3
tested with different critical geometric parameters. All specimens were manufactured from
austenitic stainless-steel of grade AISI 304. For each joint, eight unidirectional strain gauges were
placed to capture the hot spot strain perpendicular to the weld toe, and the strain parallel to the weld
toe. The maximum SCFs recorded from the experiments were then compared against the data
calculated in accordance with the design formulas of CIDECT Guideline 8 [8] and the proposed
design formulas by Tong et al. [18] and Yin et al. [19] for carbon steel hybrid tubular joints.

2. Extrapolation methods available in the literature to determine SCFs

2.1. General

The hot spot stress is also referred to as geometric stress, which results from different
deformations of tubular members, and can be influenced by the geometric shape of tubular joints.
Research on the SCFs of carbon steel tubular joints [23, 24] enabled the finding that the SCFs depend
mainly on the critical geometric parameters of β, τ and 2γ, where β=d1/b0 is the ratio of the brace
diameter and chord width, τ=t1/t0 is the ratio of brace thickness and chord thickness and 2γ=b0/t0 is the
ratio of chord width and thickness for hybrid CHS-to-SHS tubular joints.

2.2. Extrapolation methods

The hot spot stress is an important parameter, which should be determined prior to the
calculation of SCFs. However, the local stresses generated at the toe of welded region depend on
many factors, which are difficult to control in a practical engineering situation. Hence, the hot spot
stress at the weld toe cannot be measured directly; this is usually obtained by extrapolating the local
stresses from the so-called ‘extrapolation region’ that is out of the influencing region by the
geometry of the local weld toe. It should be mentioned that CIDECT Guideline 8 [8] recommends
the determination of extrapolation regions for both the CHS and SHS tubular joints. The detail is
shown in Table 1 for both the CHS and SHS tubular joints. The distance from the first measuring
point to the weld toe (Lr,min) should be greater than 0.4t or 4 mm, while the distance (Lr,max) from the

last measuring point to the weld toe should be greater than 0.65 rt or Lr,min+0.6t for the CHS

brace, and Lr,min+t for the SHS chord, where r is the radius and t is the thickness of the tubular
sections [8].

4
3. Experimental investigation

3.1. Test specimens

A total of twelve specimens including four of each T, Y and X joints were tested. For each of
the three types of joints, circular braces and square chords were used. The HOCr20Ni10Ti
stainless-steel electrode was used for the manual arc welding with an input voltage of 380V and
output voltage of 24-28V, with 180-280A current. The electrode had a groove size of 3×3 mm and a
diameter of 3.2 mm. It should be noted that as depicted in the CIDECT Guideline 8, the values of
SCFs were mainly related to the geometrical shape of the joints, and the welding process had
minimal effects on the values of SCFs. The cross-sectional dimensions of CHS and SHS tubes were
carefully designed in the fabrication of hybrid CHS-to-SHS tubular joints, which include the outer
width (b0) of the chord as 150 to 200 mm, the wall thickness (t0) of the chord as 3 to 4 mm, and the
overall length (L0) of the chord as 900 to 1200 mm. The outer diameter (d1) of the brace ranged
from 108 to 133 mm, the wall thickness (t1) of the brace being 3 mm, and the overall length (L1) of
the brace ranged from 325 to 400 mm. The chord’s overall length (L0) was selected as 6b0 to avoid
the end effects of the chord on the hot spot stress at the joint intersection region. The brace’s overall
length (L1) was designed as 3d1 to exclude the premature overall buckling of the braces [17]. Hence,
the critical geometric parameters which were varied in the experiments are brace diameter-to-chord
width ratio (β=d1/b0), brace-to-chord thickness ratio (τ=t1/t0) and, the chord width to thickness ratio
(2γ=b0/t0). The value of β was varied from 0.54 to 0.89. On the other hand, τ ranged from 0.75 to 1.00,
and the value of 2γ was 50.0, in the experiments. The specimen dimensions were carefully measured
prior to the tests, which are summarized in Table 2 with all symbols defined in Fig. 1.

3.2. Specimen labelling

The specimens were labelled in such a way that the type of hybrid tubular joints, and the
cross-sectional dimensions of SHS chord and CHS brace were defined. For example, the labels
‘T-C150×3-B108×3’ and ‘Y-C200×4-B133×3’, can be defined as follows:
• The first letters ‘T’ and ‘Y’ indicate the T and Y joints, respectively. This corresponds to
X-joint if the first letter is ‘X’.
• The chord member is referenced using the second letter ‘C’. The following phrases ‘150×3’

5
and ‘200×4’ refer to the cross-sectional dimensions of SHS chord members with 150 and 200
mm outer width (b0), along with 3 and 4 mm wall thicknesses (t0), respectively.
• The brace member is defined by the third letter ‘B’. The following phrases ‘108×3’ and
‘133×3’ represent the cross-sectional dimensions of CHS brace members with a 108 and 133
mm outer diameter (d1), and 3 mm wall thickness (t1).

3.3. Tensile coupon tests

To obtain the material properties of stainless-steel CHS and SHS, tensile coupon tests were
conducted in accordance with Chinese Code [25]. The test results are summarized in Table 3, which
includes the Young’s modulus (E0), 0.01% tensile proof stress (σ0.01), 0.2% tensile proof stress (σ0.2),
1.0% tensile proof stress (σ1.0), elongation (εf) and the ultimate tensile stress (σu). It should be noted
from the results of tensile coupon tests (σ0.2/σu ratio and εf), that the stainless-steel CHS and SHS
tubes have pronounced strain hardening and ductility properties.

3.4. Hot spot strains

In order to determine the nominal strains and control the loading eccentricity applied to the
CHS brace of hybrid tubular T, Y and X joints during the trials, four unidirectional Strain gauges
with a gauge length of 3 mm were uniformly placed along the annular direction of the brace wall at
mid-length of the CHS brace. The hot spot locations where the hot spot stress initiated, needed to be
predicted to seek the critical stress distributions at the weld toe close to the joint intersection region.
For stress distributions of carbon steel hybrid CHS-to-RHS tubular joints [22], the maximum
stresses are usually located at the typical hot spot close to the joint intersection region along 0°, 45°
and 90° to the CHS brace. Thus, a total of ten lines were chosen as the hot spot locations for all test
specimens, in which the locations along 0°, 45°, 90°, 135° and 180° to CHS brace on the SHS chord
were characterized as lines A, C, E, G and I, respectively; while the locations along 0°, 45°, 90°,
135° and 180° to CHS brace on the CHS brace were characterized as lines B, D, F, H and J,
respectively. The detailed arrangement of strain gauges is shown in Fig. 2.
The ‘extrapolation method’ (described in Section 2 of this paper) in the CIDECT Guideline 8
[8] was adopted to measure the hot spot stress at all hot spot locations. The values of Lr,min and Lr,max
were 4 and 16 mm, respectively. On the other hand, due to the unidirectional pressure perpendicular
to the weld toe, the cracks basically propagated along the weld toe region which has the greatest
6
influence on the crack propagation [26].
A total of eight unidirectional strain gauges were positioned at each hot spot, in which five
strain gauges were successively located along each line of hot spot locations to capture the hot spot
strain perpendicular to the weld toe, while another three strain gauges were positioned at the same
locations of hot spot as the first three strain gauges but perpendicular to these, respectively, to
obtain the strains parallel to the weld toe. The strain gauges were positioned to the weld toe, in
order to investigate the influence of heat affected zones resulting from welding. Furthermore, the
locations of all the strain gauges were basically within the ‘extrapolation region’ under the
recommendation of the CIDECT Guideline 8 [8]. Fig. 3 shows the strain gauge arrangements at the
hot spot region.

3.5. Testing procedure

All the specimens were axially loaded with a 50T hydraulic jack. The test setup of the hybrid
tubular T, Y and X joints stainless-steel is shown in Figs. 4a-4c. Two steel plates were welded at
both ends of the SHS chord for hybrid tubular T joints, which were attached by bolts to the
supporting plates. The supporting plates were then connected to the base supports by means of a pin,
which were firmly fixed to the strong floor through anchor bolts. Hence, the hinge boundary
condition was provided for the hybrid tubular T joints, as shown in Fig. 4a. A similar test setup was
employed for hybrid tubular Y joints, except that one end of the SHS chord was lifted to ensure that
the axial loads could be exerted vertically to the CHS brace (Fig. 4b). Both ends of the SHS chord
were unrestrained to translate and rotate in any direction without restraints for hybrid tubular X
joints. As shown in Fig. 4c, the end of the bottom CHS brace was connected to a specially designed
spherical hinge support, which could rotate in the initial loading stage to exclude the influence of
loading eccentricity and become a fixed-ended bearing at later stage of loading.
Load control method was used to apply the axial load, using the hydraulic jack. The load was
then monitored by the load cell placed between the CHS brace and the hydraulic jack. To ensure
that the hot spot stress in the joint intersection region was within the elastic range of the materials,
four unidirectional strain gauges were located uniformly along the annular direction of the brace
wall at mid-length of the CHS brace. By using the data acquisition system, readings of the load cell
and strain gauges were recorded.

7
4. Test results

4.1. Determination of SCFs

The SCFs of the tubular joints were calculated using the Eqs. 1 and 2 as shown below:

SCF =  HSS /  n (1)

4F
n = (2)
 [d − (d1 − 2t1 )2 ]
1
2

where, nominal stress (σn) was calculated from the nominal strain (εn) values obtained from the
unidirectional strain gauges positioned at mid-length of the CHS brace. It is worth noting that the
determination of the SCFs depend on the perpendicular hot spot strain (ε⊥) to the weld toe and the
parallel strain (ε//) to the weld toe. These strain values were used to obtain the SCFs, based on the
design recommendations of CIDECT Guideline 8 [8]. The strain concentration factor (SNCF) was
then calculated from the following design formula:

SNCF =  ⊥ /  n (3)

The SCFs can be calculated using the design equation recommended by Shao [27] as given
below:

∥
1 +
⊥
SCF =  SNCF (4)
1 - 2

The CIDECT Guideline 8 [8] and International Institute of Welding (IIW) [28] recommend
that the linear extrapolation method should be applied for CHS tubular joints since it has a linear
strain distribution profile at the joint intersection region. The quadratic extrapolation approach
should be applied for SHS tubular joints since it has a nonlinear distribution of strain at the joint
intersection region. Both methods of linear and quadratic extrapolations were applied in this
analysis for the hybrid CHS-to-SHS tubular joints including ‘T-C200×4-B108×3’,
‘Y-C200×4-B108×3’ and ‘X-C200×4-B108×3’, as detailed in Fig. 5. The SCFl and SCFq calculated
using linear and quadratic extrapolation methods, respectively, at the hot spot location at 90° angle
to the CHS brace, are illustrated in Table 4. The ratio of SCFq to SCFl ranged from 1.11 to 1.70.
This indicates that the stress distributions around the intersection area of the SHS chord and CHS
brace in hybrid CHS-to-SHS tubular joints are quite nonlinear. Similar nonlinear behaviour was
8
also observed by Mashiri [29] for welded thin-walled tubular joints under cyclic loads. Therefore, to
obtain the hot spot stress, the method of quadratic extrapolation was used in this study.

4.2. SCFs of hybrid tubular joints

Table 5 summarizes the SCFs of hybrid tubular T, Y and X joints. Their distributions are
plotted in Figs. 6-8, respectively. From the comparison, it can be concluded that the peak SCFs
occurred at the hot spot location at 90° angle to the CHS brace on the SHS chord that characterized
as line E for hybrid tubular T and X joints, except for the specimens ‘T-C150×3-B133×3’ and
‘X-C150×3-B133×3’. This exceptional behaviour observed in the tests for the above-mentioned test
specimens could be due to the comparatively large value of β (β=0.88), so that there was not enough
space for strain gauges to be located at line E of top flange at the SHS chord. The SCFs were
recorded from the strain gauges arranged at the side wall of the SHS chord, as detailed in Fig. 2e.
The SCF at line E of the specimen ‘Y-C150×3-B133×3’ also showed different trend in results,
similar to the specimens ‘T-C150×3-B133×3’ and ‘X-C150×3-B133×3’, as discussed above. The
maximum SCFs were also located at the hot spot locations at an angle of 90° to the CHS brace,
which characterized as lines E and F for hybrid tubular Y joints. These maximum values of SCFs at
lines E and F for CHS brace of hybrid tubular Y joints, were for smaller and larger diameters,
respectively.

4.3. Conversion coefficients between SCF and SNCF

As mentioned previously, the SCF cannot be acquired directly from measurements; this is
normally converted from the SNCF. The SNCF depends on the perpendicular hot spot strain to the
weld toe, which could be measured using the extrapolation method from the strain gauges
positioned at the hot spot location. According to the CIDECT Guideline 8 [8], it was recommended
to convert SNCF to the SCF by multiplying the conversion coefficients of 1.2 and 1.1 for CHS and
SHS tubular joints, respectively. The CIDECT Guideline 8 [8] recommend SCF=1.2×SNCF and
SCF=1.1×SNCF for CHS and SHS tubular joints, respectively. However, the current CIDECT
Guideline 8 [8] does not provide the value of conversion coefficient between the SCF and SNCF for
hybrid tubular joints.
In this paper, the conversion coefficients of SCF and SNCF for hybrid CHS-to-SHS
stainless-steel tubular joints were studied at 120 hot spot locations for all specimens, as shown in
9
Fig. 9. Based on the regression analysis of experimental results, the conversion coefficients of 1.117
and 1.263 are proposed for SHS chord and CHS brace, respectively. Hence, SCF=1.117×SNCF and
SCF=1.263×SNCF are proposed in this study for SHS chord and CHS brace of stainless-steel
hybrid CHS-to-SHS tubular joints, respectively.

5. Design rules

5.1. Design equations

The design formulas for the SCFs of tubular joints of CHS and SHS are available in CIDECT
Guideline 8 [8]. The design formulas for the T, Y and X joint SCFs are given in this section for CHS
tubular joints. The validity range of the geometric parameters given in CIDECT Guideline 8 [8] for
the design formulas of CHS and SHS tubular joints SCFs is summarized in Table 6.
However, the CIDECT Guideline 8 [8] does not provide the design formulas for the SCFs of
hybrid tubular joints, which are quite different from those tubular joints of CHS, SHS and RHS [11].
It should be noted that Tong et al. [18] and Yin et al. [19] derived design formulas for hybrid
CHS-to-SHS tubular T and Y joints, respectively. Table 6 shows the validity range of geometrical
dimensions for design formulas of SCFs of hybrid CHS-to-SHS tubular joints.

5.2. Design formulas of SCFs of CHS tubular joints

For CHS tubular joints under axial loads, the maximum SCFs generally occur at the crown or
saddle position. These crown and saddle positions are shown in Fig. 10.
For CHS chord of tubular T and Y joints [8]:

  (
SCFch_saddle,ax =   1.1  1.11 − 3  ( − 0.52)  sin1.6  + (0.8 − 6)   2  1 −  2
2
)
0.5
 sin 2 2  F2 (5)

SCFch_crown,ax =  0.2   2.65 + 5  ( − 0.65) +     (0.5 − 3) sin 


2
(6)

For CHS brace of tubular T and Y joints [8]:

 
SCFb_saddle,ax = 1.3 +   0.52  0.1  0.187 −1.25 1.1 ( − 0.96)  sin (2.7−0.01 )   F2 (7)

 
SCFb_crown,ax = 3 +  1.2  0.12  exp(−4 ) + 0.011 2 − 0.045 +    (0.2 − 1.2) (8)

For CHS chord of tubular X joints [8]:

 
SCFch_saddle,ax = 3.87       1.10 −  1.8  (sin )1.7  F2 (9)

10
 
SCFch_crown,ax =  0.2   2.65 + 5  ( − 0.65)2 − 3     sin  (10)

For CHS brace of tubular X joints [8]:

( )
SCFb_saddle,ax = 1 + 1.9    0.5   0.9  1.09 −  1.7  sin 2.5   F2 (11)


SCFb_crown,ax = 3 +  1.2  0.12  exp(−4 ) + 0.011 2 − 0.045  (12)

If   12 , F2 = 1.0 .


If   12 , F2 = 1 − (1.4 3 − 0.97  2 − 0.03)   0.04  exp − 0.7 1   −1.38   2.5  (13)

5.3. Design formulas of SCFs of SHS tubular joints

For SHS chord of tubular T and X joints [8]:

(
SCFB, ax = 0.143 − 0.204 + 0.064 2  (2 ) ) (1.377 +1.715 -1.103 )   0.75
2
(14)

(
SCFC, ax = 0.077 − 0.129 + 0.061 2 − 0.0003 2  (2 ) ) (1.565+1.874  −1.028 )  0.75
2
(15)

(
SCFD,ax = 0.208 − 0.387 + 0.209 2  (2 ) ) (0.925+ 2.389  −1.881 )  0.75
2
(16)

For SHS brace of tubular T and X joints [8]:

(
SCFA,ax = SCFE ,ax = 0.013 − 0.693 + 0.278 2  (2 ) ) (0.790+1.898 −2.109  ) 2
(17)

For tubular joints with fillet welds:


The SCFs of SHS brace should be multiplied by 1.40 for the welds at brace side.

5.4. Design formulas of SCFs of hybrid CHS-to-SHS tubular joints

For SHS chord of hybrid CHS-to-SHS tubular T joints [18]:

SCFA = (0.146 − 0.241 + 0.115 2 − 0.000803 2 ) (2 )


1.34+0.846  −0.527  2
 0.929 (18)

SCFE = (0.445 − 0.548 + 0.0927 2 ) (2 )


0.954+1.62  −0.985  2
 1.09 (19)

For CHS brace of hybrid CHS-to-SHS tubular T joints [18]:

SCFB = (0.0736 − 0.133 + 0.0688 2 − 0.00031 2 ) (2 )


0.647+2.02  −0.360  2
 −1.48+0.809 (20)

SCFF = (− 0.0156 + 0.520 − 0.470 2 ) (2 )


1.33+0.384  −0.208 2
 −0.246+0.233 (21)

For SHS chord of hybrid CHS-to-SHS tubular Y joints [19]:

( )
SCFA = − 2.966 2 + 1.242 −1.558   0.739  0.873  (sin )
0.401
(22)

11
( )
SCFC = − 4.674 2 + 4.549 − 0.120   1.364  1.195  (sin )
2.378
(23)

( )
SCFE = − 0.990 2 + 0.632 + 0.289   1.143  0.555  (sin )
1.288
(24)

For CHS brace of hybrid CHS-to-SHS tubular Y joints [19]:

( ) ( )
SCFB = − 2.032 2 + 2.170 − 0.146   1.063  − 2.548 2 + 2.734 + 0.019  (sin )
0.747
(25)

( ) ( )
SCFD = − 2.163 2 + 2.598 − 0.184   1.301  −1.836 2 + 2.507 − 0.060  (sin )
1.524
(26)

( ) ( )
SCFF = −1.203 2 + 1.242 + 0.074   1.263  − 0.771 2 + 0.247 + 0.520  (sin )
0.719
(27)

5.5. Comparison of results from experiments and the current design formulas

The SCFs at the hot spot locations of CHS brace and the SHS chords, were recorded from the
experiments and were used for comparison against the design predictions in accordance with the
CIDECT Guideline 8 [8] (Fig. 11). The experimental results for most of the hot spot locations
deviate significantly from the design values calculated in accordance with the existing design
guideline’s predictions; in particular, the design formulas for the SCFs of CHS tubular joints. This
difference in results may attribute to different geometric shapes and material properties. The main
reason for differences in results is because that the corners of SHS tubes have more abrupt change,
which will cause relatively higher value of SCFs. Also, it was found from previous studies that the
value of SCFs of CHS-to-SHS joints are between that of pure SHS-to-SHS joints and CHS-to-CHS
joints [30]. Eq. 4 showed that the material properties ν and E0 also influence the results.
It is worth noting that the fatigue behaviour of welded tubular joints depends on the maximum
value of the SCF since the hot spot stress has the largest geometrical stress located at the tubular
joint. Hence, the maximum values of the SCF recorded in the experiments were used for
comparison with those computed using the given design formulas in CIDECT Guideline 8 [8], as
summarized in Table 7. The comparison shows that the ratio of SCFExp to SCFSHS ranged from 0.10
to 0.39 for the SHS chord of hybrid tubular T, Y and X joints, with a mean value of 0.27 and a
coefficient of variation (COV) of 0.414, which means the design formulas for the SCFs of SHS
tubular joints overestimate the maximum value of SCF. Whereas, the ratio of SCFExp to SCFCHS
ranged from 1.12 to 2.14 for SHS chord of hybrid tubular Y joints with lesser diameter of CHS
brace and T joints, which means the design formulas for the SCFs of CHS tubular joints
underestimate the maximum value of SCF. The ratio of SCFExp to SCFCHS ranged from 0.37 to 0.88
12
for SHS chord of hybrid tubular Y joints with a larger diameter of CHS brace and X joints, which
means the design formulas for the SCFs of CHS tubular joints are quite unconservative for the
predictions of the peak values of SCFs. Moreover, the design formulas for the SCFs of SHS tubular
joints are unconservative for the predictions of the maximum SCFs for the CHS brace of hybrid
tubular T, Y and X joints. The SCFExp to SCFCHS ratio ranged from 3.02 to 10.23 for CHS brace of
hybrid tubular T and X joints, which means the design formulas for SCFs of CHS tubular joints
underestimate the maximum values of SCFs. Whereas, the SCFExp to SCFCHS ratio ranged from 0.67
to 1.15 for CHS brace of hybrid tubular Y joints. Therefore, it is demonstrated that the design
formulas of CIDECT Guideline 8 [8] are incapable of predicting the SCFs of CHS and SHS of
stainless-steel hybrid tubular joints.
On the other hand, the SCFs at all the hot spot locations on SHS chord and CHS brace of
stainless-steel hybrid tubular joints recorded from the experiments were also compared with the
data calculated using the design formulas derived by Tong et al. [18] and Yin et al. [19] for the
SCFs of carbon steel hybrid CHS-to-SHS tubular T and Y joints, respectively, and the comparison is
shown in Fig. 11.
The SCFs calculated using the design formulas of carbon steel hybrid CHS-to-SHS tubular
joints are close to the experimental results. Furthermore, at the locations of hot spot welds, the
maximum values of SCFs were recorded. These locations of maximum SCFs at hot spot welds were
consistent with the experimental results for all specimens, except for the specimen
‘Y-C200×4-B133×3’. The maximum value of the SCF was located at the hot spot region at an angle
of 90° to the CHS brace on either SHS chord (line E) or CHS brace (line F). However, the ratio of
SCFExp to SCFCHS-SHS ranged from 0.40 to 1.27 for SHS chord of hybrid tubular T, Y and X joints.
The mean value of the ratio of SCFExp to SCFCHS-SHS was 0.76 and a corresponding COV of 0.466
for tubular T joints. While for the case of tubular Y and X joints, the mean values of the ratio of
SCFExp to SCFCHS-SHS ranged from 0.27 to 1.38 with the corresponding COV of 0.495, which
indicates that the design predictions [18, 19] of hybrid CHS-to-SHS tubular joints significantly
deviated from the experiment results. This is because of the reason that the geometric parameters of
the test specimens are outside the scope of the validity range specified in the design formulas for the
SCFs of carbon steel hybrid CHS-to-SHS tubular joints [18, 19].

13
6. Conclusions

This paper presents an experimental investigation on SCFs of stainless-steel hybrid


CHS-to-SHS tubular T, Y and X joints. Austenitic stainless-steel (AISI 304 grade) was used to
manufacture all the test specimens. In total, twelve tests were conducted comprising T, Y and X
joints. For each joint, eight unidirectional strain gauges were placed to capture the hot spot strain
perpendicular to the weld toe and the strain parallel to the weld toe. The mechanical properties of
stainless-steel CHS and SHS tubes are reported for all test specimens. The maximum SCFs
recorded from the experiments were used for comparison with those determined using the design
formulas available in the CIDECT Guideline 8 [8] and the proposed design formulas by Tong et al.
[18] and Yin et al. [19] for carbon steel hybrid tubular joints. From the outcome of this study, the
following conclusions can be drawn:
(1) The maximum stress concentrations occurred at the hot spot locations at 90° angle to the CHS
brace on the SHS chord for hybrid tubular T and X joints, while the maximum stress concentrations
were also occurred at the hot spot locations at 90° angle to the CHS brace for hybrid tubular Y
joints. Furthermore, the conversion coefficients between the SCF and SNCF are also recommended
for both the SHS chord and CHS brace of stainless-steel hybrid CHS-to-SHS tubular joints.
(2) The SCFs recorded from the experiments at most of the hot spot locations were far from the
SCFs calculated in accordance with the design guidelines of carbon steel CHS and SHS tubular
joints, as specified in CIDECT Guideline 8 [8]. From the comparison, it was found that the carbon
steel design rules as specified in CIDECT Guideline 8 [8], are not capable of predicting the SCFs of
such stainless-steel tubular joints.
(3) The design formulas [18, 19] for carbon steel hybrid CHS-to-SHS tubular joints are better than
those of carbon steel CHS and SHS tubular joints, while predicting the SCFs of stainless-steel
hybrid CHS-to-SHS tubular joints. Besides, the locations of the maximum SCFs can also be
predicted accurately by the design formulas of carbon steel hybrid CHS-to-SHS tubular joints [18,
19].

Acknowledgements

The authors are grateful for the financial support from National Natural Science Foundation of

14
China (Grant No. 51528803), Natural Science Foundation of Guangdong Province of China (Grant
No. 2018A030313208), State Key Laboratory of Subtropical Building Science (South China
University of Technology, Grant No. 2018ZA02), and Guangdong Provincial Key Laboratory of
Durability for Marine Civil Engineering, Shenzhen Durability Center for Civil Engineering
(Shenzhen University, Grant No. GDDCE 18-5). The tests were conducted in Anhui Key Lab on
Structure and Material of Civil Engineering at Hefei University of Technology. The support
provided by the laboratory staff is gratefully acknowledged.

References

[1] Huang Y, Young B, Post-fire behaviour of ferritic stainless steel material, Constr. Build.
Mater. 2017, 157: 654–667.
[2] Dundu M, Evolution of stress–strain models of stainless steel in structural engineering
applications, Constr. Build. Mater. 2018, 165: 413–423.
[3] Shojaati.M, Beidokhti B, Characterization of AISI 304/AISI 409 stainless steel joints using
different filler materials, Constr. Build. Mater. 2017, 147: 608–615.
[4] Fan S, Jia L, Lyu X, Sun W, Chen M, Zheng J. Experimental investigation of austenitic
stainless steel material at elevated temperatures. Constr. Build. Mater. 2017, 155: 267–285.
[5] Toprac AA, Natarajan M, Erzurumlu H, Kanoo ALJ. Research in tubular joints: static and
fatigue loads. Offshore Technological Conference, Houston, USA, 1969, p. 667-680.
[6] van Wingerde AM, Packer JA, Wardenier J. New guidelines for fatigue design of HSS
connections. J. Struct. Eng. ASCE 1996, 122(2): 125-132.
[7] Feng R, Young B. Stress concentration factors of cold-formed stainless steel tubular X-joints.
J. Constr. Steel Res. 2013, 91(1): 26-41.
[8] Zhao XL, Herion S, Packer JA, Puthli RS, Sedlacek G, Wardenier J, Weynand K, van
Wingerde AM, Yeomans NF. Design Guide for Circular and Rectangular Hollow Section
Welded Joints under Fatigue Loading. Comité International pour le Développement et l’Étude
de la Construction Tubulaire (CIDECT), Verlag TÜV Rheinland, Berlin, Germany, 2001.
[9] DNVGL. Fatigue Design Offshore Steel Structures. DNVGL-RP-203, Det Norske Veritas,
Norway, 2016.

15
[10] Kuang JG, Potvin AB, Leick RD. Stress concentration in tubular joints. Offshore
Technological Conference, Houston, USA, 1975, 593-612.
[11] Gibstein MB. Parametric stress analysis of T joints. European Offshore Steels Research
Seminar, Cambridge, UK, 1978, 3-26.
[12] Gho WM, Gao F. Parametric equations for stress concentration factors in completely
overlapped tubular K(N)-joints. J. Constr. Steel Res. 2004, 60(12): 1761-1782.
[13] Woghiren CO, Brennan FP. Weld toe stress concentrations in multi-planar stiffened tubular
KK joints. International Journal of Fatigue, 2009, 31(1): 164-172.
[14] Ahmadi H, Lotfollahi-Yaghin MA, Shao YB. Chord-side SCF distribution of central brace in
internally ring-stiffened tubular KT-joints: A geometrically parametric study. Thin-Walled
Struct. 2013, 70: 93-105.
[15] Yue J, Liu Y, Zhang C, Zeng Q, Dang Z. Fatigue strength assessment on a multiplanar tubular
KK-joints by scaled model test. Journal of Offshore Mechanics and Arctic Engineering, 2015,
138(2): 1-8.
[16] Ahmadi H, Zavvar E. The effect of multi-planarity on the SCFs in offshore tubular KT-joints
subjected to in-plane and out-of-plane bending loads. Thin-Walled Struct. 2016, 106:
148-165.
[17] Mashiri FR, Zhao XL, Grundy P. Stress concentration factors and fatigue behaviour of
welded thin-walled CHS-SHS T-joints under in-plane bending. Eng. Struct., 2004, 26(13):
1861-1875.
[18] Tong LW, Zheng HZ, Mashiri FR, Zhao XL. Stress-concentration factors in circular hollow
section and square hollow section T-connections: experiments, finite-element analysis, and
formulas. J. Struct. Eng. ASCE 2013, 139(11): 1866-1881.
[19] Yin Y, Liu XF, Lei P, Zhou L. Stress concentration factor for tubular CHS-to-RHS Y-joints
under axial loads. J. Constr. Steel Res. 2018, 148: 768-778.
[20] Tagoe JB. Fatigue Performance of Hollow Section Joints-Square/Circular Cross Sections.
MSC thesis, Norwegian University of Science and Technology, Norway, 1995.
[21] Gandhi P, Berge S. Fatigue behavior of T-Joints: square chords and circular brace. J. Struct.
Eng. ASCE 1998, 124(4): 399-404.
[22] Bian LC, Lim JK. Fatigue strength and stress concentration factors of CHS-to-RHS T-joints. J.
Constr. Steel Res. 2003, 59(5): 627-640.
16
[23] Lotsberg I. On stress concentration factors for tubular Y- and T-joints in frame structures.
Marine Struct. 2011, 24(1): 60-69.
[24] van Wingerde AM, Packer JA, Wardenier J. Criteria for the fatigue assessment of hollow
structural section connections. J. Constr. Steel Res. 1995, 35(1): 71-115.
[25] Chinese Code. Metallic Materials − Tensile Testing − Part 1: Method of Test at Room
Temperature. GB/T 228.1-2010, Beijing, China, 2010. (in Chinese)
[26] Hellier AK, Connolly MP, Dover WD. Stress concentration factors for tubular Y- and T-joints.
International Journal of Fatigue, 1990, 12(1): 13-23.
[27] Shao YB. Fatigue Behaviour of Uniplanar Tubular K-joints under Axial and In-plane Bending
Loads. PhD thesis, Nanyang Technological University, Singapore, 2004.
[28] International Institute of Welding (IIW). Recommended Fatigue Design Procedure for Hollow
Section Joints. Part 1—Hot Spot Stress Method for Nodal Joints. IIW Subcommission XV-E,
IIW Doc. XV-582-85, IIW Assembly, Strasbourg, France, 1985.
[29] Mashiri FR. Thin-walled Tubular Connections under Fatigue Loading. PhD thesis, Monash
University, Australia, 2001.
[30] Packer JA, Mashiri FR, Zhao XL, Willibald S. Static and fatigue design of CHS-to-RHS
welded connections using a branch conversion method. J. Constr. Steel Res., 2007, 63(1):
82-95.

Notations

b0 Outer width of SHS chord


d1 Outer diameter of CHS brace
e Eccentricity
E0 Young’s modulus
F Axial load
Lr,max Distance from last measuring point to weld toe
Lr,min Distance from first measuring point to weld toe
L0 Overall length of chord
L1 Overall length of brace
r Radius of tubular member

17
SCFCHS SCF calculated using design formulas of CHS tubular joint
SCFCHS-SHS SCF calculated using design formulas of CHS-to-SHS tubular joint
SCFExp SCF obtained from experiment
SCFl SCF calculated using linear extrapolation method
SCFq SCF calculated using quadratic extrapolation method
SCFSHS SCF calculated using design formulas of SHS tubular joint
SCFT SCF calculated using design formulas of CHS-to-SHS tubular T-joint
SCFY SCF calculated using design formulas of CHS-to-SHS tubular Y-joint
t Wall thickness of tubular member
t0 Wall thickness of chord
t1 Wall thickness of brace
α Chord length-to-half width ratio (2L0/b0)
β Brace diameter-to-chord width ratio (d1/b0)
2γ Chord width-to-thickness ratio (b0/t0)
εf Elongation after fracture
εn Nominal strain
ε⊥ Hot spot strain perpendicular to weld toe
ε// Strain parallel to weld toe
θ Angle between circular brace and square chord
ν Poisson’s ratio
σHSS Hot spot stress
σn Nominal stress
σu Ultimate tensile stress
σ0.01 0.01% tensile proof stress
σ0.2 0.2% tensile proof stress
σ1.0 1.0% tensile proof stress
τ Brace-to-chord thickness ratio (t1/t0)

18
Table 1. Extrapolation region defined in CIDECT Guideline 8 [8] for CHS and

SHS tubular joints


Chord Brace
Distance from weld toe
saddle crown saddle crown
Lr,min ≥ 4 mm 0.4t0 0.4t1
CHS
Lr,max ≥ Lr,min+0.6t1 0.09r0 0.44 r0t 0 r1t1 0.65 r1t1

Lr,min ≥ 4 mm 0.4t0 0.4t1


SHS
Lr,max Lr,min+t0 Lr,max+t1

19
Table 2. Measured specimen dimensions of stainless-steel hybrid tubular joints
Chord (mm) Brace (mm) Geometric parameter
Specimen
b0 t0 L0 d1 t1 L1 β τ 2γ

T-C150×3-B108×3 150.23 3.014 900 108.43 3.044 325 0.72 1.01 49.84

T-C150×3-B133×3 149.99 3.022 900 132.08 3.036 400 0.88 1.00 49.63

T-C200×4-B108×3 200.15 3.957 1200 108.83 3.094 325 0.54 0.78 50.58

T-C200×4-B133×3 199.79 3.815 1200 131.23 2.884 400 0.66 0.76 52.37

Y-C150×3-B108×3 150.57 2.906 900 108.29 2.845 325 0.72 0.98 51.81

Y-C150×3-B133×3 149.80 3.000 900 131.72 3.046 400 0.88 1.02 49.93

Y-C200×4-B108×3 199.56 3.949 1200 108.39 3.044 325 0.54 0.77 50.53

Y-C200×4-B133×3 199.44 3.993 1200 132.16 3.031 400 0.66 0.76 49.95

X-C150×3-B108×3 149.93 2.960 900 108.39 3.048 325 0.72 1.03 50.65

X-C150×3-B133×3 149.81 2.966 900 131.76 3.001 400 0.88 1.01 50.51

X-C200×4-B108×3 199.90 3.971 1200 108.42 3.065 325 0.54 0.77 50.34

X-C200×4-B133×3 199.44 3.975 1200 131.44 3.016 400 0.66 0.76 50.17

20
Table 3. Material properties of stainless-steel tubes
E0 σ0.01 σ0.2 σ1.0 σu εf
Section σ0.2/σu
(MPa) (MPa) (MPa) (MPa) (MPa) (%)
□150×3 207009 353.18 430.27 457.08 780.20 57.34 0.55
□200×4 188370 299.93 423.71 442.83 758.11 54.34 0.56
○108×3 232622 370.79 441.15 458.33 750.12 54.52 0.59
○133×3 229172 366.09 433.96 448.25 753.36 63.17 0.58

21
Table 4. Determination of the SCFs of stainless-steel hybrid tubular joints
Linear Quadratic
SCFq
Specimen Position Load (kN) extrapolation extrapolation
/SCFl
SCF SCFl SCF SCFq
9.41 19.48 22.50
Chord (E) 10.00 21.53 21.24 24.11 23.71 1.12
18.24 22.70 24.51
T-C200×4-B108×3
9.41 6.99 11.89
Brace (F) 10.00 7.74 6.95 11.84 11.79 1.70
18.24 6.39 11.65
8.82 30.20 33.13
Chord (E) 14.71 28.14 28.52 31.68 31.90 1.12
19.41 27.23 30.89
Y-C200×4-B108×3
8.82 14.02 15.38
Brace (F) 14.71 13.31 12.72 15.81 15.50 1.22
19.41 10.83 15.32
5.29 13.76 14.36
Chord (E) 11.76 14.98 14.67 16.62 16.39 1.11
20.59 15.28 18.20
X-C200×4-B108×3
5.29 7.26 10.38
Brace (F) 11.76 7.13 7.43 12.21 11.36 1.53
20.59 7.89 11.49

22
Table 5. The SCFs of stainless-steel hybrid tubular joints
β= τ= 2γ= Stress concentration factor (SCF)
Specimen
d1/b0 t1/t0 b0/t0 A B C D E F G H I J
T-C150×3-B108×3 0.72 1.01 49.84 8.55 1.96 17.38 4.74 33.17 26.07 18.44 3.44 5.59 1.43
T-C150×3-B133×3 0.88 1.00 49.63 7.46 2.22 11.03 3.80 5.53 32.61 12.81 5.89 12.43 2.12
T-C200×4-B108×3 0.54 0.78 50.58 5.85 3.01 14.66 8.13 23.71 11.79 15.57 6.74 2.03 1.07
T-C200×4-B133×3 0.66 0.76 52.37 6.30 4.26 15.56 12.17 32.27 23.65 17.03 15.67 6.46 6.97

Y-C150×3-B108×3 0.72 0.98 51.81 12.43 4.06 19.68 14.42 33.24 18.96 11.11 2.08 7.97 2.36
Y-C150×3-B133×3 0.88 1.02 49.93 10.21 1.46 10.97 11.40 0.64 15.20 1.00 0.98 3.37 0.22
Y-C200×4-B108×3 0.54 0.77 50.53 14.37 2.72 17.92 9.81 31.90 15.50 14.74 7.49 4.75 3.81
Y-C200×4-B133×3 0.66 0.76 49.95 5.95 0.45 4.90 12.20 9.08 10.70 5.02 0.68 0.74 5.15
X-C150×3-B108×3 0.72 1.03 50.65 0.34 1.00 5.78 2.81 34.99 38.70 31.67 8.40 5.91 1.87
X-C150×3-B133×3 0.88 1.01 50.51 3.01 0.83 2.75 2.41 0.29 38.99 13.53 5.95 3.95 1.58

X-C200×4-B108×3 0.54 0.77 50.34 1.06 1.57 7.10 1.46 16.39 11.36 16.71 6.50 3.52 0.23
X-C200×4-B133×3 0.66 0.76 50.17 1.15 0.50 5.50 3.21 11.36 11.35 8.67 7.06 0.38 2.20

23
Table 6. Validity range of geometric parameters for the SCFs in CIDECT

Guideline 8 [8], previous studies [18, 19] and experiments


Geometric parameter β τ 2γ α θ

CHS T, Y and X joints [8] [0.2-1.0] [0.2-1.0] [15-64] [4-40] [30°-90°]

SHS T and X joints [8] [0.35-1.0] [0.25-1.0] [12.5-25] — —

CHS-to-SHS T joints [18] [0.35-1.0] [0.25-1.0] [12.5-35] — —

CHS-to-SHS Y joints [19] [0.2-0.8] [0.2-1.0] [12-40] — [30°-90°]

CHS-to-SHS T, Y and X
[0.54-0.89] [0.75-1.0] 50 12 [45°-90°]
joints in experiments

24
Table 7. Comparison of the maximum SCFs obtained from experiments, CIDECT Guideline 8 [8] and previous studies [18,

19]
Chord Brace
Maximum SCF Comparison Maximum SCF Comparison
Specimen
Experi CIDECT CIDECT SCFExp/ SCFExp/ SCFExp/ Experi CIDECT CIDECT SCFExp/ SCFExp/ SCFExp/
CHS-to-SHS CHS-to-SHS
ment CHS SHS SCFCHS SCFSHS SCFCHS-SHS ment CHS SHS SCFCHS SCFSHS SCFCHS-SHS
T-C150×3-B108×3 33.17 24.72 85.73 53.80 1.34 0.39 0.62 26.07 3.99 32.87 40.23 8.44 0.79 0.65
T-C150×3-B133×3 — 17.90 36.40 19.06 — — — 32.61 3.85 14.41 27.99 8.47 2.26 1.17
T-C200×4-B108×3 23.71 21.15 101.69 56.88 1.12 0.23 0.42 11.79 3.76 47.50 43.30 3.14 0.25 0.27
T-C200×4-B133×3 32.27 20.35 90.54 50.45 1.59 0.36 0.64 23.65 3.81 41.84 46.37 6.21 0.57 0.51
Y-C150×3-B108×3 33.24 15.50 85.73 26.10 2.14 0.39 1.27 18.96 19.82 32.87 14.33 0.96 0.58 1.32
Y-C150×3-B133×3 — 19.70 36.40 9.23 — — — 15.20 13.26 14.41 10.99 1.15 1.05 1.38
Y-C200×4-B108×3 31.90 21.83 101.69 25.99 1.46 0.31 1.23 15.50 18.82 47.50 19.76 0.82 0.33 0.78
Y-C200×4-B133×3 9.08 21.01 90.54 22.97 0.43 0.10 0.40 12.20 18.05 41.84 20.85 0.67 0.29 0.59
X-C150×3-B108×3 34.99 39.71 89.91 — 0.88 0.39 — 38.70 3.99 33.43 — 9.70 1.16 —
X-C150×3-B133×3 — 26.54 38.01 — — — — 38.99 3.81 14.62 — 10.23 2.67 —
X-C200×4-B108×3 16.39 31.19 99.76 — 0.53 0.16 — 11.36 3.75 47.23 — 3.02 0.24 —
X-C200×4-B133×3 11.36 30.51 85.73 — 0.37 0.13 — 11.35 3.76 39.86 — 3.02 0.28 —
Mean 1.10 0.27 0.76 4.65 0.87 0.83
COV 0.515 0.414 0.466 0.765 0.889 0.495

25
(a) T-joint

(b) Y-joint

(c) X-joint
Figure 1. Schematic sketch of stainless-steel hybrid tubular joints
26
I A
G C
E
(a) T-joint

I A
G C
E
(b) Y-joint

I A
G C
E
(c) X-joint

(d) Locations of strain gauge


27
(e) Locations of strain gauges at specimens with β=0.88

Figure 2. Arrangement of strain gauges

(a) T-joint (b) Y-joint (c) X-joint

Figure 3. Photos of locations of strain gauges

28
(a) T-joint

Hydraulic jack
Load cell

Hinge

Anchor
Base

(b) Y-joint

(c) X-joint
Figure 4. Test setup of stainless-steel hybrid tubular joints
29
Figure 5. Extrapolation methods

30
(a) Chord (b) Brace

Figure 6. SCF curves of stainless-steel hybrid tubular T-joints

(a) Chord (b) Brace

Figure 7. SCF curves of stainless-steel hybrid tubular Y-joints

(a) Chord (b) Brace

Figure 8. SCF curves of stainless-steel hybrid tubular X-joints


31
(a) SHS chord

(b) CHS brace

Figure 9. Conversion coefficients between SCF and SNCF

Figure 10. The definition of the crown and saddle positions


32
(a) T-C150×3-B108×3

(b) T-C150×3-B133×3

(c) T-C200×4-B108×3
33
(d) T-C200×4-B133×3

(e) Y-C150×3-B108×3

(f) Y-C150×3-B133×3
34
(g) Y-C200×4-B108×3

(h) Y-C200×4-B133×3

(i) X-C150×3-B108×3
35
(j) X-C150×3-B133×3

(k) X-C200×4-B108×3

(l) X-C200×4-B133×3
Figure 11. Comparison of the SCFs obtained from experiments, CIDECT
Guideline 8 [8] and previous studies [18, 19]
36

View publication stats

You might also like