You are on page 1of 43

Gardner, L. and Yun, X. (2018) Description of stress-strain curves for cold-formed steels.

Construction and Building Material, 189, 527-538.

Description of stress-strain curves for cold-formed steels

Leroy Gardner and Xiang Yun*

Department of Civil and Environmental Engineering, Imperial College London, London, SW7 2AZ,
UK.
* Corresponding author. E-mail address: x.yun14@imperial.ac.uk (X. Yun)

Graphical abstract:
1.8
0.6
 f 
0.5
 u  0.61  y  1.6
 fu 
1.4
0.4
1.2

fu/fy
0.3 1.0
εu

0.2 Stress f 0.8


f u / f y  1  130 f y 
1.4
0.1
E0.2 0.6

0.0
fu 0.4
0 200 400 600 800 1000 1200 1400
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
fy/fu fy m
fy (N/mm2)
f  fy  f  fy   f  fy  6
    u   0.2  u     0.2
20
E0.2  E0.2   f u  f y 
5

15 4
mtest

n 3
10
 f 
ntest

f
  0.002  
E  fy  2
m  1  3.3
fy
 
ln 4 
5 fu
n
ln  f y  0.05  1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
0 E fy/fu
0 5 10 15 20
npred
ε0.2 εu Strain ε

Abstract: Cold-formed steels are generally characterized by a rounded stress-strain response

with no sharply defined yield point. It is shown herein that this material behaviour can be

accurately described by a two-stage Ramberg-Osgood model provided that the values of the

key input parameters can be established. The focus of the present paper is to develop predictive

expressions for these key parameters to enable the full engineering stress-strain response of

cold-formed steels to be represented. The predictive expressions are based on the analysis of a

comprehensive set of material stress-strain data collected from the literature. In total, more than
1
700 experimentally-derived stress-strain curves on cold-formed steel material have been

collected from around the world, covering a range of steel grades, thicknesses and cross-section

types. The strength enhancement in the corner regions of cold-formed sections has also been

analysed and the applicability of existing predictive models has been evaluated. Finally,

standardized values of strain-hardening exponents used in the Ramberg-Osgood model have

been recommended for both flat and corner material in cold-formed steel sections. The

proposed stress-strain curves are suitable for use in advanced numerical simulations and

parametric studies on cold-formed steel elements.

Keywords: Cold-formed steels; Constitutive modelling; Material modelling; Ramberg-

Osgood model; Stress-strain curve; Strain hardening; Strength enhancement.

1. Introduction

Cold-formed steel members are widely used in building construction due to their high strength

and stiffness-to-weight ratios and ease of prefabrication and mass production. Press-braking

and roll-forming are the two primary methods employed in the manufacture of cold-formed

steel products. In both methods, steel sheet or strip, which would typically have been cold-

rolled, serves as the feed material. In press-braking, sheet material is formed into the desired

geometric shape by applying predetermined bends along its length, which are generally

accomplished by punch and die sets. The press-braking process is commonly used to produce

cold-formed open sections, such as angles and channels. Roll-forming is a process by which

steel sheet or strip is passed through a series of rollers that progressively shape the material

into the final cross-section profile. Different levels of cold-work (plastic deformation) are

generated during the manufacturing process, resulting in changes to the stress-strain

characteristics of the material. Generally, cold-work results in a more rounded stress-strain

2
response with an increased yield strength and, to a lesser extent, an increased ultimate strength,

but reduced ductility. Non-uniformity of material properties is also commonly seen in cold-

formed sections due to the varying level of cold-work experienced at different locations around

the section shape. The corner regions of cold-formed cross-sections undergo large plastic

deformation from cold bending due to their tight corner radii, hence resulting in a higher degree

of strength enhancement but a corresponding loss in ductility.

Constitutive modelling is an essential part of structural engineering and a key component of

analytical, numerical and design models. A number of material models have been developed

to describe the nonlinear stress-strain response of metallic materials over the last few decades,

as reviewed in the next section, the most widely used of which are based on the general

expression proposed by Ramberg and Osgood [1] and later modified by Hill [2]. A number of

studies have focused on the material modelling of stainless steels (Rasmussen [3], Quach et al.

[4], Arrayago et al. [5] and Tao and Rasmussen [6]), but values and predictive expressions for

the key parameters in these models require modification for applicability to the particular

stress-strain characteristics of cold-formed steels. The aim of the present study is to develop an

accurate stress-strain model for use in the advanced simulation of cold-formed steel members,

based upon and validated against data from over 700 experimental stress-strain curves collected

from the global literature, covering a wide range of steel grades, production routes and cross-

section types. A rounded stress-strain response is associated with structural cross-sections that

experience sufficient cold-work during sheet/section forming to erode the yield plateau, and is

particularly common for cold-formed tubular sections. Cold-formed steels that exhibit this

characteristic rounded stress-strain response with no yield plateau are the focus of the present

paper.

3
2. Overview of existing models and previous work

2.1. Existing stress-strain models

The Ramberg-Osgood model (Ramberg and Osgood [1] and Hill [2]) is widely used to describe

the rounded stress-strain response of aluminum alloys, stainless steels and cold-formed carbon

steels. The Ramberg-Osgood model (see Eq. (1)) is generally defined using three basic

parameters – the Young’s modulus E, the material yield strength fy (taken as the 0.2% proof

stress) and the strain hardening exponent n. The strain hardening exponent n can be determined

using an additional proof stress e.g. the 0.01% proof stress σ0.01, as given by Eq. (2), or the

0.05% proof stress σ0.05, as given by Eq. (3). The latter has been shown by Rasmussen and

Hancock [7] and Arrayago et al. [5] to yield more consistent n values in comparison to curves

fitted to stainless steel stress-strain data by regression analysis.

n
f  f 
   0.002  (1)
E  fy 
 

ln20
n
ln  f y  0.01 
(2)

ln 4 
n
ln  f y  0.05 
(3)

Eq. (1) has been shown to be capable of accurately representing different regions of the stress-

strain curve of stainless steel, depending on the choice of the n parameter, but found to be

generally incapable of accurately representing the full stress-strain curve with a single value of

n. This observation led to a number of developments and extensions to the Ramberg-Osgood

model. Mirambell and Real [8] proposed a smooth two-stage stress-strain model, utilising the
4
conventional Ramberg-Osgood relationship (Eq. (1)) up to the yield strength and introducing

a second Ramberg-Osgood curve commencing at the yield strength fy and continuing up to the

ultimate tensile strength fu, as given by Eq. (4). Continuity of position and slope is achieved at

the transition point (i.e. the yield strength fy) between the two stages (i.e., between Eq. (1) and

Eq. (4)). Fig. 1 shows a typical cold-formed steel stress-strain curve, together with the key

material parameters used in the two-stage Ramberg-Osgood model.

m
f  fy  f  fy   f  fy 
    u   0.2  u     0.2 , for f y  f  f u (4)
E0.2  E0.2   f u  f y 

In Eq. (4), εu is the strain at the ultimate strength fu, ε0.2 is the total strain at the yield strength fy

(0.2% proof stress), E0.2 is the tangent modulus at the 0.2% proof stress, given by Eq. (5), and

m is the strain hardening exponent for the second part of the two-stage model, which can be

determined from the ultimate strength and an intermediate strength. Note that Eq. (1) remains

applicable for strengths less than or equal to the yield strength fy.

E
E0.2  (5)
E
1  0.002n
fy

The two-stage Ramberg-Osgood model was simplified by Rasmussen [3] by approximating

fu  fy
the plastic strain term  u   0.2  in Eq. (4) to the ultimate strain εu, as expressed in Eq.
E0.2

(6). In addition, Rasmussen [3] also developed predictive expressions for the second strain

hardening exponent m, the ultimate strain εu and ultimate strength fu for stainless steel alloys,

as given by Eqs. (7)-(9), respectively, using the three basic Ramberg-Osgood parameters (E, fy

5
and n). As a result, only the three basic parameters are required to describe full stress-strain

curves for different stainless steel grades. Note that the simplified expression [Eq. (6)] predicts

a stress-strain curve that does not pass exactly through the ultimate strength point (εu, fu),

though the discrepancies due to the approximation of ultimate strain in the Rasmussen model

are small especially for the more ductile stainless steel grades (austenitic and duplex), as

highlighted in Arrayago et al. [5]. The Rasmussen model, along with the predictive expressions

for the key material parameters [Eqs. (7) and (8)], has been included in Annex C of EN 1993-

1-4 [9]. Further recent studies (Afshan et al. [10], Real et al. [11], Arrayago et al. [5] and Tao

and Rasmussen [6]) have been carried out to assess the accuracy of the predictive equations

(Eqs. (7)-(9)), and revised expressions that provide more accurate parameter predictions have

been proposed where necessary.

m
f  fy  f  fy 
  u    0.2 , for f y  f  f u
 f u  f y 
(6)
E0.2  

fy
m  1  3.5 (7)
fu

fy
u  1 (8)
fu

 fy
 0.2  185 E , for austenitic and duplex stainless steels

fy 
 (9)
fu  f
0.2  185 y
 E
1  0.0375(n  5) , for all stainless steel alloys

6
Since an ultimate strength in compression cannot be obtained due to the absence of the necking

phenomenon, it was proposed by Gardner and Nethercot [12] to use the 1% proof stress σ1.0 in

place of the ultimate strength fu in Eq. (4), leading to Eq. (10) for strengths greater than the

yield strength fy.

n0.2,1.0
f  fy    fy   f  fy 
   1.0   0.2  1.0     0.2 , for f y  f  f u (10)
E0.2  E0.2   σ1.0  f y 

where ε1.0 is the total strain at the 1% proof stress σ1.0, and n0.2,1.0 is a strain hardening exponent

representing a curve that passes through the yield strength point (ε0.2, fy) and the 1% proof stress

point (ε1.0, σ1.0). This approach was found to provide an excellent match to the experimental

stress-strain data, in both tension and compression, up to strains of approximately 10% [12],

which covers the maximum strains experienced in general structural applications.

For certain scenarios in which large plastic strains are likely to be encountered, such as the

simulation of cold-forming processes, connections and structures under extreme loads, an

accurate prediction of stress-strain curves up to very high strains is often needed. A three-stage

full-range stress-strain model, which utilised Eq. (10) up to ε2.0 (i.e., the total strain at the 2%

proof stress σ2.0), followed by a third stage based on the assumption that a linear relationship

exists between the true stress and the nominal strain over the strain range of ε2.0 < ε ≤ εu, was

proposed by Quach et al. [4] for stainless steels. The model is accurate, but the predictive

expressions for the additional material parameters are rather lengthy and their scope of

application is limited. Hradil et al. [13] developed a generalized multistage model based on the

Ramberg-Osgood expression (i.e., using the same concept as the Mirambell and Real model),

which enables a nonlinear stress-strain curve to be more accurately described by dividing the

7
curve into multiple segments, though increasing the number of segments increases the number

of material parameters and hence the complexity of the model.

All the aforementioned material models are based upon and calibrated against stainless steel

stress-strain curves, but studies into stress-strain models for cold-formed steels are relatively

scarce. Abdel-Rahman and Sivakumaran [14] proposed an idealized elastoplastic multi-linear

stress-strain model to describe the gradual yielding and strain hardening behaviour of cold-

formed steel materials. Quach and Huang [15] extended the applicability of the Mirambell and

Real model (see Eq. (4)) to cold-formed steels and proposed two alternative predictive

equations for the second strain hardening exponent m, which ensure that the second stage of

the stress-strain curve passes either through the 1% or 2% proof stress, as given by Eqs. (11)

and (12). Note that for coupons that exhibit very low values of εu, the second strain hardening

exponent m may not be obtained using the 1% or 2% proof stress; in this scenario, the value of

m can be conservatively taken as unity [15]. However, the accuracy of the resulting stress-

strain curves that utilise the strain hardening exponent m from Eqs. (11) or (12) depends largely

on the ability to predict the selected intermediate proof stress, and although predictive

expressions for σ1.0 and σ2.0 have been proposed by Quach and Huang [15], the scatter of the

collected experimental data points for σ1.0 and σ2.0 relative to the predictions is significant. The

proposed equations for m (Eqs. (11) and (12)) are thus more suitable for use where the

measured 1% or 2% proof stress is known.

   f y  1.0  f y   fu  fy 
ln  0.008  1.0    ln   u   0.2   f  fy
 E E0.2   E0.2    (11)
m , for  u  max   0.2  u , 1.0 
 
ln  1.0  f y  ln( f u  f y )  E0.2 

8
   f y  2.0  f y   fu  fy 
ln  0.018  2.0    ln   u   0.2   f  fy
 E E0.2   E0.2    (12)
m , for  u  max   0.2  u ,  2.0 
 
ln  2.0  f y  ln( f u  f y )  E0.2 

A comparative study of various material models used to describe the nonlinear behaviour of

stainless steels has been conducted by Real et at. [11], which highlighted that the three-stage

Ramberg-Osgood models yield the best fit to experimental stress-strain curves at high strains,

although more material parameters are required for their definition. It was also pointed out that

the optimized values (i.e., those found by fitting to measured stress-stain curves) of the first

and second strain hardening exponents obtained for different two-stage Ramberg-Osgood

models (Mirambell and Real [8], Rasmussen [3] and Gardner and Nethercot [12]) are similar.

Considering the balance between accuracy and practicality, it is proposed that the Mirambell

and Real expressions (i.e., Eqs. (1) and (4)) be adopted as the basis of the stress-strain curves

for cold-formed steels developed herein. This model was favoured over the Rasmussen model

 fu  fy  fu  fy
since while the approximation that  u   u    0.2   (i.e., that the  0.2  term is
 E0.2  E0.2

small in comparison with εu) is reasonable for austenitic and duplex stainless steels that are

very ductile, the errors become more significant for cold-formed steels, which have

considerably lower ductility. Values and predictive expressions for the key input parameters

are derived in the remainder of this paper following analysis of an extensive set of experimental

stress-strain curves.

2.2. Existing models for predicting strength enhancements in cold-formed steel sections

Over the past few decades, there have been a number of studies aimed at developing suitable

models for predicting the enhanced yield strength in the corner regions of cold-formed steel

9
and stainless steel cross-sections that arise as a result of plastic deformation experienced during

production, a review of which has been made by Rossi et al. [16]. Early investigations into the

strength enhancement in the corner regions of cold-formed steel sections were carried out by

Karren [17], who conducted a series of coupon tests on virgin, flat and corner materials and

proposed a semi-empirical model to predict the increased yield strength in the corners fy,c, as

given in the following equations:

Bc
f y,c  fy (13)
 ri t 

in which

2
 f   f 
Bc  3.69  u   0.819  u   1.79 (14)
 fy   fy 
   

and

 fu 
  0.192   0.068
 f y 
(15)
 

Karren’s model assumes that the strength enhancement in the corner regions of cold-formed

steel sections is dependent on (i) the ratio of the ultimate tensile strength (fu) to the yield

strength (fy) of the unformed (virgin) material, which is indicative of the potential for cold-

working, and (ii) the ratio of the inner corner radius (ri) to the thickness of the steel sheet (t),

which is indicative of the induced level of plastic strain. Note that values of fy and fu are

provided in mill certificates but, for design purposes, the values given in material specifications

should be adopted. Karren’s model is included in the AISI Specification for the Design of Cold-

formed Steel Structural Members [18], referred to hereinafter as the “AISI Specification”.

Abdel-Rahman and Sivakumaran [14] observed, based on tensile coupon test results on cold-

10
formed steel channel sections, that strength enhancements also exist in areas adjacent to the

corners though these are not as significant as in the curved corner areas themselves. Following

their findings, a modification factor to Eq. (13) was proposed to give the average enhanced

yield strength within the corner zone, which was defined as the curved corner area plus adjacent

flat regions either side of the corner extending to a distance of πri/2. Similarly, Gardner et al.

[19] modified the predictive model given in the AISI Specification [Eqs. (13)-(15)] based on

test results on corner material extracted from cold-formed steel box sections, and the revised

values of coefficients are given in Eqs. (16) and (17). The modified predictive model was found

to provide more accurate and consistent corner yield strength predictions than those obtained

from the AISI Specification for cold-formed square and rectangular hollow sections.

2
 f   f 
Bc  2.9  u   0.752  u   1.09 (16)
 fy   fy 
   

 fu 
  0.23   0.041
 f y 
(17)
 

An alternative model to predict the strength enhancement in the corner regions of cold-formed

sections based on the determination of the plastic strains associated with the dominant stages

in the fabrication process and the adoption of an inverted compound Ramberg-Osgood stress-

strain relationship [20] for the unformed material was proposed by Rossi [21]. This predictive

model was established without introducing any empirical parameters and was thus deemed

applicable to any structural sections formed from non-linear materials. More recently, Rossi et

al. [16] proposed a modified and simplified approach featuring a simple power law model in

place of the inverted compound Ramberg-Osgood model, to represent the stress-strain response

11
of the unformed material, and a through thickness averaged plastic strain, in place of the

maximum surface plastic strain adopted by Rossi [21]. The Rossi et al. [16] model is given by

Eqs. (18) and (19) for the determination of the enhanced yield strength for the flat faces fy,f of

cold-rolled box sections and the corner regions fy,c of cold-rolled or press-braked sections,

respectively.

f y,f  p  f,av   0.2  , but  f u


q
(18)

f y,c  p  c,av   0.2  , but  f u


q
(19)

where εf,av and εc,av refer to the through thickness averaged plastic strain induced during the

forming of the flat regions of cold-rolled box sections and the corner regions of press-braked

and cold-rolled sections, as defined by Eqs. (20) and (21), respectively, and p and q are material

parameters calculated directly from the basic properties of the unformed material, as given in

Eqs. (22) and (23). In Eqs. (20) and (21), b and h are the width and depth, respectively, of the

cold-rolled box-sections, and t is the thickness. Note that the 0.85 factor included in the Rossi

et al. [16] model for reliability purposes has been excluded for the comparisons made in the

present study in order to directly assess the accuracy of the predictive expressions.

 t   t 
 f,av      (20)
 900  2b  h  2t  

t
 c,av  (21)
2(2ri  t )

12
fy
p (22)
 0.2
q

ln( f y / f u )
q (23)
ln( 0.2 /  u )

3. Experimental database and data processing method

3.1. Experimental database

In this section, the collection and analysis of experimental data to establish predictive

expressions for the key material parameters used in Eqs. (1) and (4) for cold-formed steels are

described. The database comprised results from over 700 tensile coupon tests gathered from a

broad spectrum of existing experimental programmes conducted around the world. A summary

of the references for the test data, the number of experimental results and the steel grades is

provided in Table 1. The tested coupons were cut either from cold-formed steel sheets or the

flat/corner regions of a range of cold-formed steel sections in the direction parallel to the rolling

direction, including square hollow sections (SHS), rectangular hollow sections (RHS),

elliptical hollow sections (EHS), semi-oval hollow sections (SOHS), circular hollow sections

(CHS), polygonal hollow sections (PHS), lipped and plain channel sections, cassette sections,

angle sections, Zee-sections and C-shaped sections. The collated data covers steel grades of

nominal yield strength between 235 N/mm2 and 1100 N/mm2. In total, 729 tensile coupon test

results (581 flat coupons and 148 corner coupons) from the published literature were assembled

to establish the predictive expression for the ultimate strength fu, of which 598 results were

used to evaluate the predictive model for the ultimate strain εu, since the test values of εu were

not explicitly reported for 131 of the 729 tests. A total of 356 full-range stress-strain curves

were reported and analyzed to develop suitable predictive expressions for the strain hardening

exponents n and m employed in the two-stage Ramberg-Osgood model (Mirambell and Real
13
[8]) for cold-formed steels. Note that full details of the measurement methods were not reported

in all studies, but where stated, strain gauges were generally employed to record the initial part

of the stress-strain curve up to the yield strength, while extensometer readings were used for

the remainder of the curve. A range of gauge lengths were adopted in the different studies

considered, but this would only be expected to influence the stress-strain curve after the

occurrence of necking (i.e. beyond the ultimate tensile strength), which is not within the scope

of this investigation. Also note that all the analysed stress-strain curves displayed the

characteristic rounded shape. Stress-strain curves for the flat faces of some press-braked

sections exhibited a yield plateau that is typical of hot-rolled steels, due to the limited amount

of cold-work experienced during the fabrication processes, but these results were excluded

from the database.

3.2. Experimental data processing

The collected stress-strain curves were carefully analysed to extract the values of the key

parameters. Firstly, the Young’s modulus E (i.e., the slope of the linear portion of the stress-

strain curves) was determined. Following analysis of the collated experimental results, it is

recommended that the Young’s modulus is derived over a stress interval between 0.2fy and

0.4fy, to avoid the influence of any test data anomalies at the beginning of the stress-strain

curves and any nonlinearity at higher stress levels. A similar stress interval was suggested by

Huang and Young [69] for the analysis of cold-formed carbon steel (G450) stress-strain curves.

It should be noted that since fy (i.e., the 0.2% proof stress) depends on E, an initial value of the

Young’s modulus, taken as 200,000 N/mm2, was assumed for the purposes of identifying the

0.2-0.4fy stress interval. The average slope over the stress interval was then taken as the

Young’s modulus. Having determined the value of E, the proof stresses, including the 0.2%

proof stress which is conventionally taken as the yield strength, corresponding to a plastic strain

14
i σi were obtained as the points of intersection between the measured stress-strain curve and a

shifted line from the origin to the plastic strain i with the same slope as the Young’s modulus.

The ultimate strength fu and the corresponding ultimate strain εu were also captured.

To determine the strain hardening exponents n and m from the measured tensile coupon test

results, the least squares regression method, where the sum of the squared residuals between

the experimental curves and the considered material model is minimised, was employed. Since

the loading rates applied throughout tensile coupon tests are generally not constant, with a

lower loading rate applied in the initial region to obtain sufficient data points for determining

the Young’s modulus and higher loading rates thereafter to avoid excessive testing times, the

data points along the measured stress-strain curves are not evenly distributed and sometimes

exhibit some disturbance (e.g., a small jump in the stress-strain curve when the loading rate

increases). The uneven distribution and any discontinuities in the measured data points may

result in some inaccuracy in the regression analysis, and thus should be eliminated. As an

additional measure to ensure the strain hardening exponents were accurately captured, a curve

fitting approach was employed prior to the least squares regression analysis, as described below,

in order to represent the experimental stress-strain curves with an evenly and smoothly

distributed set of data points.

In accordance with Mirambell and Real’s two-stage Ramberg-Osgood model, the experimental

stress-strain curves were first divided into two segments for strains below and above ε0.2. Each

of the two segments was found to be accurately represented by a 7th order polynomial

(Sadowski et al. [70] and Yun and Gardner [71]), as given by Eqs. (24) and (25), where a1 to

a7 and b1 to b7 are two sets of trial coefficients to be determined. All the polynomial fits to the

experimental data points exhibited a coefficient of determination R2 ≥ 0.96, indicating an

15
accurate representation of the measured stress-strain curves. Evenly distributed data points

could then be obtained from the fitted polynomial and the process of least squares regression

could then be carried out to determine the strain hardening exponents n and m.

7
f     a k k , for 0     0.2 (24)
k 1

7
f    f y   bk (   0.2 ) k , for  0.2     u (25)
k 1

The collated experimental results were categorized into two groups: flat coupons and corner

coupons. The curved coupons extracted from cold-formed CHS exhibited similar stress-strain

behaviour to those cut from the flat faces of cold-formed sections, and were thus grouped with

the flat coupons. On the basis of the collated test data for cold-formed steels, recommended

values and predictive expressions for the key material parameters used in the two-stage

Ramberg-Osgood model are provided in the following section.

4. Analysis of results and recommendations

4.1. Young’s modulus

The average Young’s modulus value obtained from the data assembled in this study for cold-

formed steels was 203,000 N/mm2. This value is the same as that recommended in AISI [18]

but is slightly lower than the value 210,000 N/mm2 given in EN 1993-1-1 [62].

4.2. First strain hardening exponent n

The first strain hardening exponent n can be calculated using the classic expression proposed

by Ramberg and Osgood [1], as given by Eq. (2) and adopted in EN 1993-1-4 for the modelling
16
of stainless steel stress-strain behaviour. The expression requires the analytical curve to pass

through the 0.01% and the 0.2% proof stresses. In more recent studies (Arrayago et al. [5],

Mirambell and Real [8], Rasmussen and Hancock [7] and Real et al. [11]), it has been proposed

that use of the 0.05% proof stress, as given by Eq. (3), instead of the 0.01% proof stress

provides more accurate predictions of n for stainless steels.

Assessment of the two expressions (Eqs. (2) and (3)) for cold-formed steels is presented in Fig.

2, where the test values ntest, obtained from the experimental stress-strain curves through the

described data processing method, are plotted against the predicted values npred using Eqs. (2)

and (3). It can be clearly seen that Eq. (3) provides more accurate and consistent predictions of

the test values ntest than Eq. (2). Key statistical values, including the mean and coefficient of

variation (COV) of the predicted-to-test results (npred/ntest), calculated using either Eq. (2) or

Eq. (3), are presented in Table 2; the numerical comparisons also indicate the improved

accuracy of Eq. (3). It is therefore recommended that the calculation of n for cold-formed steels

be based on the 0.05% proof stress, and that the 0.05% proof stress values are routinely

measured and reported by researchers in future experimental programs.

Aside from providing a predictive expression for the determination of n from measured data,

it is also desirable to have representative numerical values for n that capture the general degree

of roundedness of the first stage of the stress-strain curve for cold-formed steels when measured

data are not available. The recommended values, taken as the average of the collated ntest values,

are provided in Table 3, with a value of 7.6 for flat coupons and a slightly lower value of 7.0

for corner coupons. Since there are insufficient data to enable a meaningful evaluation of the

influence of loading direction (transverse or longitudinal) and loading sense (compression and

17
bending) on the value of n, and the influence of steel grade has been found to be small, it is

proposed that no distinction is made between the above parameters in assigning the value of n.

4.3. Second strain hardening exponent m

When measured stress-strain data are available, analogous to Eq. (2) and (3) for n, the value of

m can be determined by imposing that the second stage of the stress-strain curve passes through

a defined intermediate point. The suitability of using the two alternative points proposed by

Quach and Huang [15], (ε1.0, σ1.0) or (ε2.0, σ2.0), leading to Eqs. (11) and (12) respectively for

the definition of m, is assessed against the mtest values determined using the described data

processing method. Note that for coupons that exhibit very low ductility, i.e., very low εu values,

m cannot be obtained using the defined intermediate points and a conservative value of unity

was thus recommended by Quach and Huang [15], resulting in a straight line from (ε0.2, fy) to

(εu, fu); measured stress-strain curves in this category were excluded from the assessment

performed herein, resulting in 196 stress-strain curves available for evaluating the accuracy of

the two expressions (Eqs. (11) and (12)). Assessment of the two expressions for the

determination of m is presented in Fig. 3, where the test values mtest are plotted against the

predicted values mpred using Eq. (11) or Eq. (12). In addition, Table 4 presents the statistical

results concerning the mean and COV values of mpred, using the different predictive expressions,

relative to the test values mtest. As illustrated in Fig. 3 and Table 4, Eq. (11) offers considerably

more accurate predictions of the mtest values than Eq. (12) and it is therefore recommended that

Eq. (11) can be used for the determination of m, provided that σ1.0 can be accurately measured

from the stress-strain curve.

When measured stress-strain data are not available, it is also desirable to have a predictive

expression and representative values for m. Eq. (7), expressed in terms of the ratio of yield

18
strength to ultimate strength fy/fu, is provided in Annex C of EN 1993-1-4 for the second strain

hardening exponent m of stainless steels. Observing a similar trend in the cold-formed steel

data assembled herein, a linear relationship between m and fy/fu ratios was established. The test

data mtest are plotted against the corresponding fy/fu ratios for both flat and corner coupons in

Fig. 4. It may be observed that the test data from the flat coupons and the corner coupons follow

a similar trend, though the corner coupons generally exhibit higher yield-to-ultimate strength

ratios fy/fu due to the higher degree of cold-work that they would have experienced, and higher

values of mtest. By linear regression, a revised expression was proposed for cold-formed steels,

as given in Eq. (26). The most recently proposed predictive expression for the second strain

hardening exponent m for stainless steels [5] has also been plotted in Fig. 4 for comparison

purposes. The predictive expression for stainless steels underestimates the values of m for the

majority of cold-formed steel coupon tests, as shown in Fig. 4, while the new proposal (Eq.

(26)) provides good average predictions, with a mean ratio of the predicted-to-test values

mpred/mtest of 0.99 and a low COV of 0.09, as summarized in Table 4. In instances where the

measured fu is unavailable, it is recommended to adopt the predicted value from Eq. (30), as

presented in Section 4.5. The average values of mtest for both flat and corner coupons are also

given in Table 3.

fy
m  1  3.3 (26)
fu

4.4. Ultimate strain εu

For the prediction of ultimate strain for both hot-rolled and cold-formed steels, with an

additional lower limit imposed for hot-rolled steels, Eq. (27) was proposed by Yun and Gardner

[71]. This equation was also proposed by Bock et al. [72] for ferritic stainless steel. Benefiting

from a larger database of results, the accuracy of Eq. (27) has been further assessed herein, as
19
shown in Fig. 5, where the experimental ultimate strains εu are plotted against the

corresponding fy/fu ratios, together with the predictive expression (Eq. (27)). Fig. 5 shows good

agreement between Eq. (27) and the flat coupon data, with the mean predicted-to-test values

εu,pred/εu,test being 0.97, and a moderate COV of 0.38.

 f 
 u  0.61  y  (27)
 fu 

For the corner coupons, the test data generally follow a less inclined slope in Fig. 5 due to their

lower measured ductility. However, it should be noted that additional challenges exist in the

testing of curved corner coupons, such as the unavoidable non-uniform cold-work through the

thickness of the coupons, the problem of measuring the cross-sectional area accurately and the

difficulty of applying the tensile force through the centroid of the curved coupons in order to

avoid any possible bending effects; these challenges affect the measurement of εu and other

parameters. Karren [17] proposed the use of extra-long coupons for corner tensile tests in order

to minimize bending and flattening of the corner at the mid-length of the specimens during

testing, while additional recommendations for the testing of curved coupons were also made

by Gardner and Nethercot [12] and Huang and Young [69]. It is thus considered necessary to

conduct further tensile coupon tests on corner specimens, carefully following the recommended

test procedures, to obtain more reliable data to underpin the assessment of proposed equations

for corner coupons.

4.5. Ultimate tensile strength fu

In this section, the collated test results for cold-formed steels have been used to assess the

variation of the fu/fy ratio with yield strength fy, as shown in Fig. 6(a). The corresponding results

20
from 537 hot-rolled carbon steel tensile coupon tests, collected by Yun and Gardner [71], have

also been plotted in Fig. 6(a) for comparison purposes. The general trend of decreasing ratios

of fu/fy with increasing yield strength fy may be clearly observed, as seen in previous studies

(Fukumoto [73], Langenberg [74] and Wang et al. [75]). The empirical relationship between

fu/fy and fy proposed by Fukumoto [73] and Langenberg [74], as given by Eqs. (28) and (29),

respectively, are also depicted in Fig. 6(a). Both expressions may be seen to follow the trend

of the data for the hot-rolled steels fairly closely, but tend to overestimate the values of fu/fy for

cold-formed steels; this is also highlighted in Table 5 where the key statistical results of the

predicted-to-test values of fu/fy are tabulated. The discrepancies are such that separate

expressions to predict the relationship between fu/fy and fy for hot-rolled and cold-formed steels

are deemed necessary.

f u / f y  0.83  203.8 / f y (28)


f u / f y  1  0.72e 
0.0027f y  1
(29)

On the basis of regression analysis, a new equation to describe the variation of fu/fy with yield

strength fy for cold-formed steels is proposed, as given by Eq. (30). The accuracy of this

proposal is assessed against the assembled cold-formed steel coupon test data, as presented in

Fig. 6(b). As indicated in Table 5, Eq. (30) yields an improved mean value of the predicted-to-

test ratio for fu/fy equal to unity and a reduced COV relative to Eqs. (28) and (29) of 0.04,

confirming the accuracy of Eq. (30) for application to cold-formed steels.

f u / f y  1  130 f y  , for cold-formed steels


1.4
(30)

21
Inspired by the simplicity and accuracy of Eq. (30), a revised expression using the same

mathematical form as Eq. (30), but with different coefficients selected based on regression

analysis, is also proposed for hot-rolled carbon steels, as given by Eq. (31). Note that Eq. (31)

is based upon hot-rolled steels with yield strength ranging from 200 N/mm2 to 1000 N/mm2

and thus considered to be applicable to these steel grades; the corresponding range of

applicability of Eq. (30) based on the collected test data is 340 N/mm2 to 1200 N/mm2. The

proposed expression (Eq. (31)) offers similar accuracy and consistency as the two existing

expressions (Eqs. (28) and (29)) for hot-rolled carbon steels, as shown in Fig. 6(c) and Table

5, but avoids the anomaly of Eq. (28) of having predicted values of fu/fy that are less than unity

for high values of yield strength fy and is simpler than Eq. (29).

f u / f y  1   200 f y 
1.75
, for hot-rolled steels (31)

Eqs. (30) and (31) may be used to predict the ultimate tensile strength of cold-formed and hot-

rolled steels, respectively, in instances where the yield strength is known (e.g. by measurement)

but the corresponding ultimate tensile strength is not. This predicted value of fu may then be

used for the determination of the other key parameters (m and εu) that rely on knowledge of the

ultimate tensile strength in their predictive formulae (Eqs. (26) and (27), respectively).

4.6. Corner yield strength fy,c

The accuracy of the three different predictive models described earlier for the determination of

the enhanced yield strength of corner materials extracted from cold-formed steel sections is

assessed herein. Since not all required parameters were reported for the experimental data

collected from the literature (see Table 1), the predictive models could only be evaluated
22
against a sub-set of the database. As a result, a total of 121 sets of test data (Gardner et al. [19],

Key et al. [30], Kyvelou et al. [34], Kyvelou et al. [35], Salmi et al. [48] and Zhao and Hancock

[60]), with each set containing test results from a corner coupon and a corresponding flat

coupon cut from either the same cold-formed steel section or the virgin/sheet material, was

selected and used in the assessment. Overall, it is observed that the corner yield strength may

be significantly higher than that of the base material, with an average enhancement in yield

strength of around 25% for the data assembled herein. Table 6 compares the corner yield

strengths determined from the experiments fy,c,test with those obtained using the three different

predictive models fy,c,pred, presented in terms of the predicted-to-test ratios. For some of the

collected test results, the properties of the unformed (virgin) material were not available; in

these instances, the values of fy and fu used in the predictive models were instead taken as the

measured yield and ultimate strengths of the flat material cut from the corresponding sections.

Note that the strength enhancement in the flat faces of cold-formed steel sections is typically

relatively small (or even zero in the case of press-braked sections) due to the limited amount

of cold-work induced during the fabrication process, though can be quite substantial in cold-

rolled box-sections, as observed by Gardner et al. [19] and Guo et al. [76]. As indicated in

Table 6, the predictive models proposed by Rossi et al. [16] and Gardner et al. [19] give

similarly accurate predictions of fy,c,test, in terms of both the mean and the COV of fy,c,pred/fy,c,test,

while the model set out in the AISI Specification [18] generally over-estimates fy,c,test and shows

a higher scatter (COV). Though the two former methods provide similar accuracy, the Rossi el

al. [16] model is more universally applicable and is thus recommended for predicting the corner

yield strength of cold-formed steel sections. Note that, for design purposes, it is tentatively

proposed that the 0.85 factor recommended by Rossi et al. [16] for stainless steel also be applied

to cold-formed carbon steel pending further investigation. This requires conducting reliability

23
analysis against test results on cold-formed steel structural cross-sections. The ultimate tensile

strength of the corner material fu,c may then be approximated from Eq. (30) with fy taken as fy,c.

5. Comparisons of proposed model with measured stress-strain curves

Representative comparisons between four experimental stress-strain curves and the

corresponding predicted curves from the proposed stress-strain models are shown in Fig. 7.

The accuracy of three different predicted curves, which are generated based on: Case (1) the

measured material parameters of E, fy, fu, εu, n and m, as reported in Table 7; Case (2) the

measured material parameters of E, fy and fu, the recommended value of n as given in Table 3,

and the predicted values of εu and m using Eqs. (27) and (26), respectively; Case (3) the

measured value of only fy, the average values of E (230,000 N/mm2) and n (see Table 3), and

the predicted values of m, εu and fu obtained from Eqs. (26), (27) and (30) respectively, are

assessed in this section. The comparisons show that the proposed two-stage Ramberg-Osgood

model generally provides an accurate representation of the experimental cold-formed steel

stress-strain curves over the full range of the response, with greater accuracy achieved with

increased knowledge of the input parameters i.e. moving from Case (3) to Case (1). For the

typical case where the values of E, fy and fu are known and the values of εu, n and m are predicted

i.e. Case (2), the modelled stress-strain curves are consistently in very close agreement with

the experimental stress-strain curves.

6. Summary of proposals

Based on the described analyses, the proposed approach, i.e. adoption of the two-stage

Ramberg-Osgood model originally developed for stainless steel [8], but with the revised

predictive expressions detailed in Section 4 for describing the material stress-strain behaviour

of cold-formed steels, is summarized in this section. The stress-strain relationship is given by:

24
 f  f 
n

  0.002   , for f  f y
 E  fy 
 
  m
(32)
 f  fy  fu  fy   f  fy 
    u   0.2      0.2 , for f y  f  f u
 E0.2  E0.2   f u  f y 

The recommended values and predictive expressions for the key input parameters into Eq. (32)

are as follows:

The first strain hardening exponent n may be determined from [5]:

ln 4 
n
ln  f y  0.05 
(Eq. (3))

when the measured 0.05% proof stress is available, or otherwise taken from Table 3.

The tangent modulus of the stress-strain curve at the yield strength (0.2% proof stress) E0.2 is

defined as:

E
E0.2  (Eq. (5))
E
1  0.002n
fy

The strain εu corresponding to the ultimate tensile strength fu may be obtained from:

 fy 
 u  0.61   (Eq. (27))
 f u 

25
The ultimate tensile strength fu, if unknown, may be determined using the following predictive

expression:

f u / f y  1  130 f y 
1.4
(Eq. (30))

The second strain hardening exponent m may be determined either from Eq. (26) or Eq. (11)

[15], the latter of which requires knowledge of the measured 1% proof stress σ1.0:

fy
m  1  3.3 (Eq. (26))
fu

   f y  1.0  f y   fu  fy 
ln  0.008  1.0    ln   u   0.2   f  fy
 
m 
E E0.2   E0.2 
, for  u  max   0.2  u , 1.0 
 
ln  1.0  f y  ln( f u  f y )  E0.2 

(Eq. (11))

7. Conclusions

A comprehensive study into the constitutive modelling of cold-formed steels has been

presented. The two-stage Ramberg-Osgood model developed by Mirambell and Real [8] for

stainless steels is proposed herein to describe the engineering stress-strain relationship for cold-

formed steels. Predictive expressions and numerical values for the required parameters to be

input into the model have been devised based on the analysis of a large set of experimental

stress-strain data collected from the literature. For the prediction of corner material properties,

the proposals of Rossi et al [16] are recommended. Using the proposals, full stress-strain curves

for cold-formed steels can be established based on only the yield strength, though greater

accuracy is achieved when more input information is provided; typically, the three input

parameters of E, fy and fu would be known. The predicted stress-strain curves are shown to be
26
in excellent agreement with test results, and the proposals made herein are considered to be

appropriate for use in design by advanced computational analysis, as well as numerical

parametric studies [77].

Acknowledgements

The financial support from the China Scholarship Council (CSC) for the second author’s PhD

study at Imperial College London is gratefully acknowledged.

References

[1] W. Ramberg, W.R. Osgood, Description of stress-strain curves by three parameters,

Technical Note No. 902, National Advisory Committee for Aeronautics, Washington, D.C.,

USA, 1943.

[2] H.N. Hill, Determination of stress-strain relations from the offset yield strength values,

Technical Note No. 927, National Advisory Committee for Aeronautics, Washington, D.C.,

USA, 1944.

[3] K.J.R. Rasmussen, Full-range stress-strain curves for stainless steel alloys, J. Constr. Steel

Res. 59 (1) (2003) 47-61.

[4] W.M. Quach, J.G. Teng, K.F. Chung, Three-stage full-range stress-strain model for

stainless steels, J. Struct. Eng. ASCE 134 (9) (2008) 1518-1527.

[5] I. Arrayago, E. Real, L. Gardner, Description of stress-strain curves for stainless steel alloys,

Mater. Des. 87 (2015) 540-552.

[6] Z. Tao, K.J.R. Rasmussen, Stress-strain model for ferritic stainless steels, J. Mater. Civil.

Eng. ASCE 28 (2) (2015) 06015009.

[7] K.J.R. Rasmussen, G.J. Hancock, Design of cold-formed stainless steel tubular members.

II: Beams, J. Struct. Eng. ASCE 119 (8) (1993) 2368-2386.

27
[8] E. Mirambell, E. Real, On the calculation of deflections in structural stainless steel beams:

an experimental and numerical investigation, J. Constr. Steel Res. 54 (1) (2000) 109-133.

[9] EN 1993-1-4, Eurocode 3: Design of steel structures − Part 1-4: General rules –

Supplementary rules for stainless steels, European Committee for Standardization (CEN),

Brussels, Belgium, 2006.

[10] S. Afshan, B. Rossi, L. Gardner, Strength enhancements in cold-formed structural sections

− Part I: Material testing, J. Constr. Steel Res. 83 (2013) 177-188.

[11] E. Real, I. Arrayago, E. Mirambell, R. Westeel, Comparative study of analytical

expressions for the modelling of stainless steel behaviour, Thin-Walled Struct. 83 (2014) 2-11.

[12] L. Gardner, D.A. Nethercot, Experiments on stainless steel hollow sections − Part 1:

Material and cross-sectional behaviour, J. Constr. Steel Res. 60 (9) (2004) 1291-1318.

[13] P. Hradil, A. Talja, E. Real, E. Mirambell, B. Rossi, Generalized multistage mechanical

model for nonlinear metallic materials, Thin-Walled Struct. 63 (2013) 63-69.

[14] N. Abdel-Rahman, K.S. Sivakumaran, Material properties models for analysis of cold-

formed steel members, J. Struct. Eng. ASCE123 (9) (1997) 1135-1143.

[15] W.M. Quach, J.F. Huang, Stress-strain models for light gauge steels, Procedia Eng. 14

(2011) 288-296.

[16] B. Rossi, S. Afshan, L. Gardner, Strength enhancements in cold-formed structural sections

− Part II: Predictive models, J. Constr. Steel Res. 83 (2013) 189-196.

[17] K. Karren, Corner properties of cold-formed shapes, J. Struct. Div. ASCE 93 (1) (1967)

401-433.

[18] AISI S100-2007, North American Specification for the Design of Cold-formed Steel

Structural Members, American Iron and Steel Institute (AISI), 2007.

[19] L. Gardner, N. Saari, F. Wang, Comparative experimental study of hot-rolled and cold-

formed rectangular hollow sections, Thin-Walled Struct. 48 (7) (2010) 495-507.

28
[20] K. Abdella, Inversion of a full-range stress–strain relation for stainless steel alloys, Int J

Non Linear Mech. 41 (3) (2006) 456-463.

[21] B. Rossi, Mechanical properties, residual stresses and structural behaviour of thin-walled

stainless steel profiles (Ph.D. thesis), Univ. of Liège, Belgium, 2008.

[22] J. Becque, T. Wilkinson, The capacity of grade C450 cold-formed rectangular hollow

section T and X connections: An experimental investigation, J. Constr. Steel Res. 133 (2017)

345-359.

[23] M.T. Chen, B. Young, Tests of cold-formed steel elliptical hollow section beams,

Proceedings of the 7th International Conference on Coupled Instabilities in Metal Structures,

Baltimore, Maryland, USA, 2016.

[24] M.T. Chen, B. Young, Experimental investigation on flexural behaviour of cold-formed

steel semi-oval hollow sections, Proceedings of the 8th International Conference on Steel and

Aluminium Structures, Hong Kong, China, 2016.

[25] M. D'Aniello, R. Landolfo, V. Piluso, G. Rizzano, Ultimate behaviour of steel beams

under non-uniform bending, J. Constr. Steel Res. 78 (2012) 144-158.

[26] D.C. Fratamico, S. Torabian, X.L. Zhao, K.J.R. Rasmussen, B.W. Schafer, Experimental

study on the composite action in sheathed and bare built-up cold-formed steel columns, Thin-

Walled Struct. 127 (2018) 290-305.

[27] C. Hui, Moment redistribution in cold-formed steel purlin systems (Ph.D. thesis), Imperial

College London, London, UK, 2014.

[28] H. Jiao, X.L. Zhao, T. Ghanbari Ghazijahani, Capacity of high strength square hollow

section steel tube filled with concrete and hardwood timber, Proceedings of the 16th

International Symposium on Tubular Structures, Melbourne, Australia, 4-6 December, 2017.

[29] Y.B. Kwon, G.J. Hancock, Tests of cold-formed channels with local and distortional

buckling, J. Struct. Eng. ASCE 118 (7) (1992) 1786-1803.

29
[30] P.W. Key, S.W. Hasan, G.J. Hancock, Column behaviour of cold-formed hollow sections,

J. Struct. Eng. ASCE 114 (2) (1988) 390-407.

[31] J. Kesti, Local and distortional buckling of perforated steel wall studs (Ph.D. thesis),

Helsinki Univ. of Technology, Finland, 2000.

[32] O. Kaitila, Web crippling of cold-formed thin-walled steel cassettes (Ph.D. thesis),

Helsinki Univ. of Technology, Finland, 2004.

[33] M. Kettler, Elastic-plastic cross-sectional resistance of semi-compact H- and hollow

sections (Ph.D. thesis), Graz Univ. of Technology, Austria, 2008.

[34] P. Kyvelou, L. Gardner, D.A. Nethercot, Composite action between cold-formed steel

beams and wood-based floorboards, Int. J. Struct. Stab. Dyn. 15 (8) (2015) 1540029-1540046.

[35] P. Kyvelou, L. Gardner, D.A. Nethercot, Testing and analysis of Composite Cold-Formed

Steel and Wood−Based Flooring Systems, J. Struct. Eng. ASCE 143 (11) (2017) 04017146.

[36] G.W. Li, Y.Q. Li, Overall stability behaviour of axially compressed cold-formed thick-

walled steel tubes, Thin-Walled Struct. 125 (1) (2018) 234-244.

[37] H.T. Li, B. Young, Material properties of cold-formed high strength steel at elevated

temperatures, Thin-Walled Struct. 115 (2017) 289-299.

[38] H.T. Li, B. Young, Tests of cold-formed high strength steel tubular sections undergoing

web crippling, Eng. Struct. 141 (2017) 571-583.

[39] D. Liu, H. Liu, Z. Chen, X. Liao, Structural behaviour of extreme thick-walled cold-

formed square steel columns, J. Constr. Steel Res. 128 (2017) 371-379.

[40] J.L. Ma, Behaviour and design of cold-formed high strength steel tubular members (Ph.D.

thesis), Univ. of Hong Kong, Hong Kong, China, 2016.

[41] J.L. Ma, T.M. Chan, B. Young, Material properties and residual stresses of cold-formed

high strength steel hollow sections, J. Constr. Steel Res. 109 (2015) 152-165.

30
[42] W.F. Maia, L.C.M. Vieira Jr, B.W. Schafer, M. Malite, Experimental and numerical

investigation of cold-formed steel double angle members under compression, J. Constr. Steel

Res. 121 (2016) 398-412.

[43] M. Mohan, T. Wilkinson, Identification of parameters for continuum damage mechanics

of ductile damage evolution of cold formed Grade C450 steel rectangular hollow sections,

Proceedings of the 16th International Symposium on Tubular Structures, Melbourne, Australia,

4-6 December, 2017.

[44] F. McCann, L. Gardner, S. Kirk, Elevated temperature material properties of cold-formed

steel hollow sections, Thin-Walled Struct. 90 (2015) 84-94.

[45] J. Nseir, Development of a new design method for the cross-section capacity of steel

hollow sections (Ph.D. thesis), Univ. of Liège, Liège, Belgium, 2015.

[46] K.J.R. Rasmussen, Bifurcation of locally buckled point symmetric columns −

Experimental investigations, Thin-Walled Struct. 44 (11) (2006) 1175-1184.

[47] N. Schillo, M. Theofanous, L. Gardner, M. Feldmann, Material properties and local

buckling behaviour of high strength steel hollow sections, Proceedings of the 7th European

Conference on Steel and Composite Structures, Napoli, Italy, 2014.

[48] P. Salmi, J. Kouhi, R. Puthli, S. Herion, O. Fleischer, F. Espiga, P. Croce, E. Bayo, R.

Goni, T. Björk, R. Ilvonen, Design rules for cold-formed structural hollow sections, EUR

21973, European Commission, 2006.

[49] L. Sundararajah, M. Mahendran, P. Keerthan, New design rules for lipped channel beams

subject to web crippling under two-flange load cases, Thin-Walled Struct. 119 (2017) 421-437.

[50] T.G. Singh, K.D. Singh, Structural performance of YSt-310 cold-formed tubular steel stub

columns, Thin-Walled Struct. 121 (2017) 25-40.

31
[51] A.T. Tran, M. Veljkovic, C. Rebelo, L.S. da Silva, Resistance of cold-formed high strength

steel circular and polygonal sections − Part 1: Experimental investigations, J. Constr. Steel Res.

120 (2016) 245-257.

[52] S. Torabian, D.C. Fratamico, B.W. Schafer, Experimental response of cold-formed steel

Zee-section beam-columns, Thin-Walled Struct. 98 (2016) 496-517.

[53] A. Uzzaman, J.B. Lim, D. Nash, J. Rhodes, B. Young, Web crippling behaviour of cold-

formed steel channel sections with offset web holes subjected to interior-two-flange loading,

Thin-Walled Struct. 50 (1) (2012) 76-86.

[54] T.J. Wilkinson, The plastic behaviour of cold-formed rectangular hollow section (Ph.D.

thesis), Univ. of Sydney, Australia, 1999.

[55] B. Young, Tests and design of fixed-ended cold-formed steel plain angle columns, J. Struct.

Eng. ASCE 130 (12) (2004) 1931-1940.

[56] B. Young, Experimental investigation of cold-formed steel lipped angle concentrically

loaded compression members, J. Struct. Eng. ASCE 131 (9) (2005) 1390-1396.

[57] J. Yan, B. Young, Column tests of cold-formed steel channels with complex stiffeners, J.

Struct. Eng. ASCE 128 (6) (2002) 737-745.

[58] B. Young, K.J.R. Rasmussen, Behaviour of cold-formed singly symmetric columns, Thin-

Walled Struct. 33 (2) (1999) 83-102.

[59] S. Yi, B. Young, Experimental investigation of concrete-filled cold-formed steel elliptical

stub columns, Proceedings of the 16th International Symposium on Tubular Structures,

Melbourne, Australia, 4-6 December, 2017.

[60] X.L. Zhao, G.J. Hancock, Square and rectangular hollow sections subject to combined

actions, J. Struct. Eng. ASCE 118 (3) (1992) 648-667.

[61] A. Zhu, X. Zhang, H. Zhu, J. Zhu, Y. Lu, Experimental study of concrete filled cold-

formed steel tubular stub columns, J. Constr. Steel Res. 134 (2017) 17-27.

32
[62] EN 1993-1-1, Eurocode 3: Design of steel structures − Part 1-1: General rules and rules

for buildings, European Committee for Standardization (CEN), Brussels, Belgium, 2005.

[63] AS 1163, Structural steel hollow sections, Standards Australia, Sydney, Australia, 1991.

[64] AS 1397, Steel sheet and strip − Hot-dip zinc-coated or aluminium/zinc-coated, Standards

Australia, Sydney, Australia, 1993.

[65] EN 10027-1, Designation systems for steels – Part 1: Steel names, European Committee

for Standardization (CEN), Brussels, Belgium, 2005.

[66] GB 50017-2003, Code for design of steel structure, Beijing, China, 2003 (in Chinese).

[67] EN 1993-1-12, Eurocode 3: Design of steel structures − Part 1-12: Additional rules for the

extension of EN 1993 up to steel grades S 700, European Committee for Standardization (CEN),

Brussels, Belgium, 2007.

[68] Indian Standard 4923, Hollow steel sections for structural use − Specification, Bureau of

Indian Standards, New Delhi, 1997.

[69] Y. Huang, B. Young, The art of coupon tests, J. Constr. Steel Res. 96 (2014) 159-175.

[70] A.J. Sadowski, J.M. Rotter, T. Reinke, T. Ummenhofer, Statistical analysis of the material

properties of selected structural carbon steels, Structural Safety 53 (2015) 26-35.

[71] X. Yun, L. Gardner, Stress-strain curves for hot-rolled steels, J. Constr. Steel Res. 133

(2017) 36-46.

[72] M. Bock, L. Gardner, E. Real, Material and local buckling response of ferritic stainless

steel sections, Thin-Walled Struct. 89 (2015) 131-141.

[73] Y. Fukumoto, New constructional steels and structural stability, Eng. Struct. 18 (10) (1996)

786-791.

[74] P. Langenberg, Relation between design safety and Y/T ratio in application of welded

high strength structural steels, International Symposium on Applications of High Strength

33
Steels in Modern Constructions and Bridges-Relationship of Design Specifications, Safety and

Y/T Ratio, Beijing, China, 2008, 28-46.

[75] J. Wang, S. Afshan, N. Schillo, M. Theofanous, M. Feldmann, L. Gardner, Material

properties and compressive local buckling response of high strength steel square and

rectangular hollow sections, Eng. Struct. 130 (2017) 297-315.

[76] Y.J. Guo, A.Z. Zhu, Y.L. Pi, F. Tin-Loi, Experimental study on compressive strengths of

thick-walled cold-formed sections, J. Constr. Steel Res. 63 (5) (2007) 718-723.

[77] X. Yun, L. Gardner, The continuous strength method for the design of cold-formed steel

non-slender tubular cross-sections, Eng. Struct. 175 (2018) 549-564.

34
Tables:

Table 1. Summary of number and details of cold-formed steel coupon test data.

Material parameters fy, fu


References Steel grade Full stress-strain curves
and/or εu
Afshan et al. [10] S355a 12 (SHS/RHS) 8 (SHS/RHS)
b
Becque and Wilkinson [22] C450 24 (SHS/RHS) -
Chen and Young [23] - 8 (EHS) 8 (EHS)
Chen and Young [24] - 12 (SOHS) 12 (SOHS)
D'Aniello et al. [25] - 12 (SHS/RHS) -
Fratamico et al [26] - 8 (Lipped and plain channels) -
a
Gardner et al. [19] S235 10 (SHS/RHS) 10 (SHS/RHS)
Hui [27] 390 MPa 7 (C-shaped sections) -
Jiao et al. [28] 760 MPa 2 (SHS) 2 (SHS)
c
Kwon and Hancock [29] G500 10 (Lipped channels) -
Key et al. [30] 350 MPa 22 (SHS) -
d
Kesti [31] S350 2 (Lipped channels) -
d
Kaitila [32] S350 32 (Cassette sections) 32 (Cassette-sections)
a a
Kettler [33] S275 /S355 24 (SHS/RHS) 24 (SHS/RHS)
Kyvelou et al. [34] S450a 6 (C-shaped sections) 6 (C-shaped sections)
Kyvelou et al. [35] S450a 9 (C-shaped sections) 9 (C-shaped sections)
e
Li and Li [36] Q345 12 (SHS/RHS) -
f
Li and Young [37] S700 /S900 3 (SHS/RHS) -
f
Li and Young [38] S700 /S900 11 (SHS/RHS) -
e
Liu et al. [39] Q235 9 (SHS) -
S700f/S900/
Ma [40]; Ma et al. [41] 22 (SHS/RHS/CHS) 22 (SHS/RHS/CHS)
S1100
Maia et al. [42] - 1 (Angle section) -
b
Mohan and Wilkinson [43] C450 22 (SHS/RHS) -
a
McCann et al. [44] S355 3 (CHS/SHS/RHS) -
a
Nseir [45] S355 17 (SHS/RHS/CHS) 16 (SHS/RHS)
c
Rasmussen [46] G500 1 (Sheet) -
f f
Schillo et al. [47] S500 /S700 34 (SHS/RHS) 19 (SHS/RHS)
Salmi et al. [48] S355a/S460a 215 (SHS/RHS) 112 (SHS/RHS)
G450c/G500c/
Sundararajah et al. [49] 10 (Lipped channels) -
G550c
Singh and Singh [50] Yst-310g 21 (SHS/RHS) -
f
Tran et al. [51] S650 18 (CHS/PHS) -
Torabian et al. [52] 550 MPa 12 (Zee-sections) -
Uzzaman et al. [53] - 5 (Lipped channels) -
b b
Wilkinson [54] C450 /C350 76 (RHS) 76 (RHS)
c c
Young [55] G450 /G500 3 (Plain angle sections) -
c c
Young [56] G450 /G500 3 (Lipped angle sections) -
c
Yan and Young [57] G450 1 (Lipped channel) -

35
Table 1. Summary of number and details of cold-formed steel coupon test data (Continued).

Material parameters fy, fu


References Steel grade Full stress-strain curves
and/or εu
Young and Rasmussen [58] G450c 4 (Lipped and plain channels) -
Yi and Young [59] 363-530 MPa 8 (EHS) -
Zhao and Hancock [60] - 10 (SHS/RHS) -
235 MPa and
Zhu et al. [61] 8 (SHS and sheets) -
530 MPa
Total 729 356
a: Specified according to EN 1993-1-1 [62]
b: Specified according to Australian Standard AS 1163 [63]
c: Specified according to Australian Standard AS 1397 [64]
d: Specified according to EN 10027-1 [65]
e: Specified according to GB 50017-2003 [66]
f: Specified according to EN 1993-1-12 [67]
g: Specified according to Indian Standard 4923 [68]

Table 2. Statistical results of npred/ntest using different predictive expressions when measured σ0.01 and σ0.05 are
available.
npred/ntest
Eq. (2) Eq. (3)
Mean 0.90 0.98
COV 0.20 0.09

Table 3. Recommended representative values for strain hardening exponents n and m.

n m
Flat coupons 7.6 3.8
Corner coupons 7.0 4.2

Table 4. Statistical results of mpred/mtest using different predictive expressions.

mpred/mtest
Measured σ1.0 and σ2.0 are available Measured fy/fu is available
Eq. (11) Eq. (12) Eq. (26)
Mean 0.98 1.37 0.99
COV 0.17 0.41 0.09

36
Table 5. Statistical results of (fu/fy)pred/(fu/fy)test sing different predictive expressions.

(fu/fy)pred/(fu/fy)test
Fukumoto [72] Langenberg [73] Proposal in current study
Eq. (30) for cold-formed steels
Eq. (28) Eq. (29)
and Eq. (31) for hot-rolled steels
Cold-formed Mean 1.07 1.07 1.00
steels COV 0.05 0.05 0.04
Hot-rolled Mean 1.02 1.01 1.00
steels COV 0.07 0.07 0.07
Mean 1.04 1.03 1.00
All
COV 0.07 0.07 0.06

Table 6. Statistical results of fy,c,pred/fy,c,test using different predictive models.

fy,c,pred/fy,c,test
AISI Specification [18] Gardner et al. [19] Rossi et al. [16]
Eqs. (12)-(14) Eqs. (12), (15) and (16) Eqs. (17), (18), (20)-(22)
Mean 1.14 0.99 1.00
COV 0.08 0.07 0.07

Table 7. Measured material parameters of cold-formed steels used for comparison.

E fy fu εu
Reference Steel grade Tensile test specimen n m
(N/mm2) (N/mm2) (N/mm2) (%)
Nseir [45] S355 RHS 220 × 120 × 6-CF 206,600 430 529 14.8 6.8 4.4
Wilkinson [54] C450 TS03D 150 × 50 × 3.0 201,000 450 530 6.1 15.8 3.1
Ma [40] S700 H120 × 120 × 4 194,000 707 813 5.7 7.6 5.3
Ma [40] S900 V89 × 4 211,000 1020 1074 2.0 7.8 4.1

37
Figures:

Fig. 1. Typical cold-formed steel stress-strain curve with definitions of key material parameters.

18

16

14

12

10
ntest

4
Eq. (2)
2
Eq. (3)
0
0 2 4 6 8 10 12 14 16 18
npred

Fig. 2. Evaluation of predictive expressions for strain hardening exponent n for cold-formed steels
when measured σ0.01 and σ0.05 are available.

38
8

5
mtest

1 Eq. (11)
Eq. (12)
0
0 1 2 3 4 5 6 7 8
mpred

Fig. 3. Evaluation of predictive expressions for strain hardening exponent m for cold-formed steels
when measured σ1.0 and σ2.0 are available.

4
mtest

Eq. (26)
3

2
m=1+2.8fy/fu [5] Flat coupons
Corner coupons
1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
fy/fu

Fig. 4. Evaluation of predictive expressions for strain hardening exponent m for cold-formed steels
when measured ratios of fy/fu are available.

39
0.6
Flat coupons
0.5 Corner coupons

0.4

0.3
εu

Eq. (27)
0.2

0.1

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
fy/fu

Fig. 5. Assessment of ultimate strain εu for cold-formed steels.

2.0
Hot-rolled steel
Eq. (28) Cold-formed steel
1.8

1.6

Eq. (29)
1.4
fu/fy

1.2

1.0

0.8

0.6
0 200 400 600 800 1000 1200 1400
fy (N/mm2)

(a)

40
2.0
Cold-formed steel
1.8

1.6 Eq. (30)

1.4
fu/fy

1.2

1.0

0.8

0.6
0 200 400 600 800 1000 1200 1400
fy (N/mm2)

(b)

2.0
Hot-rolled steel
1.8
Eq. (31)
Eq. (28)
1.6
Eq. (29)
1.4
fu/fy

1.2

1.0

0.8

0.6
0 200 400 600 800 1000 1200 1400
fy (N/mm2)

(c)
Fig. 6. Assessment of the variation of fu/fy with yield stress fy for (a) both cold-formed and hot-rolled
steels, (b) cold-formed steels only and (c) hot-rolled steels only.

41
(a)

(b)

42
(c)

(d)
Fig. 7. Comparison of the predicted stress-strain curves based on different measured material
parameters with test stress-strain curves of (a) S355 flat coupon (Nseir [45]), (b) C450 flat coupon
(Wilkinson [54]), (c) S700 flat coupon (Ma [40]) and (d) S900 curved coupon (Ma [40]).

43

You might also like