You are on page 1of 9

Journal of Materials Processing Technology 170 (2005) 536–544

Artificial neural networks for modelling the mechanical properties


of steels in various applications
Z. Sterjovski ∗ , D. Nolan, K.R. Carpenter, D.P. Dunne, J. Norrish
Faculty of Engineering, University of Wollongong, Northfields Avenue, NSW 2522, Australia

Received 16 November 2004; received in revised form 26 November 2004; accepted 26 May 2005

Abstract

The application of artificial neural networks (ANNs) in predicting some key properties of steels is discussed in detail. This paper reports on
the effectiveness of three back-propagation artificial neural network models that predict (i) the impact toughness of quenched and tempered
pressure vessel steel exposed to multiple postweld heat treatment (PWHT) cycles, (ii) the hardness of the simulated heat affected zone in
pipeline and tap fitting steels after in-service welding and (iii) the hot ductility and hot strength of various microalloyed steels over the
temperature range for strand or slab straightening in the continuous casting process. Predicted and actual experimental values for each model
are well matched and highlight the success of applying ANNs in predicting mechanical properties. The capability of ANNs in predicting
multiple outputs (hot ductility and hot strength) is also demonstrated.
The sensitivity, which is a measure of the response of an output across the range of an individual input variable, of key input variables
(individual alloys and/or process steps) for each model is shown to be in agreement with findings of both the experimental investigation and
reports in the literature. Although this paper shows that ANNs can be employed for optimizing steel and process design parameters, some
difficulty can arise when inter-relationships exist between input variables. An understanding of the inter-relationships between input variables
is essential for interpreting the sensitivity data and optimizing design parameters. It is argued that artificial neural network models can be
developed that have the capacity to eliminate the need for expensive experimental investigation in areas, such as welding (new and repair),
inspection and testing, and manufacturing processes.
© 2005 Elsevier B.V. All rights reserved.

Keywords: Artificial neural network modelling; Steels; Impact toughness; Hardness; Hot ductility; Continuous casting; In-service welding

1. Introduction in analysing the tensile properties of welds in power plant


steels [3], the effect of carbon content on the hot strength of
Artificial neural networks (ANNs) are computational net- steels [4] and the impact toughness of welds based on com-
works that attempt to simulate the processes that occur in mon welding parameters [5].
the human brain and nervous system during pattern recog- This paper introduces three different back-propagation
nition, information filtering and functional control [1]. Alter neural network models which can predict the (i) impact
[2] describes ANNs as systems that recognize patterns based toughness of quenched and tempered steels exposed to var-
on the data used to train them, where each training case is ious postweld heat treatment (PWHT) cycles, (ii) simulated
characterized by a set of inputs and a result (output or multi- heat affected zone toughness of pipeline steels resulting from
ple outputs). The use of these networks to make predictions in-service welding and (iii) hot ductility (reduction of area
of the mechanical properties of materials is a relatively new (ROA)) and hot tensile strength of microalloyed steels. A
concept, but one that has received considerable interest in back-propagation neural network is the most common type of
recent years. ANNs have been reported to be very effective neural network and it undergoes its learning phase by calcu-
lating an error between the predicted and actual output. This
∗ Corresponding author. Tel.: +61 2 42214842. ANN incorporates a hidden layer that is used to establish the
E-mail address: zoran@uow.edu.au (Z. Sterjovski). inter-relationships between the input variables and their rela-

0924-0136/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2005.05.040
Z. Sterjovski et al. / Journal of Materials Processing Technology 170 (2005) 536–544 537

tionship with the output to minimize the error between the the straightening operation in the continuous casting process.
actual and predicted output. This model is designed to predict two outputs, and in doing so
The testing of QT steels to qualify them for pressure vessel it provides an insight of the functional behaviour of artificial
use is a time consuming and expensive process. Hence, the neural networks in cases of multiple outputs.
development of a reliable model to predict relevant mechan- As well as predicting specific mechanical properties as a
ical properties, such as impact toughness as a function of function of process and material parameters, all three models
alloy content (steel design) and heat treatment conditions allow the optimization of these parameters through a sensitiv-
(process parameters), is a worthwhile goal for the pressure ity analysis of the input variables. Sensitivity is a measure of
vessel industry. Modelling impact toughness can be difficult the response of output across the entire range of an individual
due to the well-known scatter that is associated with Charpy input variable [6]. In short, the aim of this paper is to underline
V notch impact testing. However, ANNs automate the mod- the usefulness and effectiveness of artificial neural networks
elling process and this improves the possibility of success. to predict, with relatively high accuracy, various properties
Accurate prediction of hardness of the heat affected zone of steels and discuss the advantages and limitations of ANNs
of a weldment as a function of composition and cooling time in steel and process design.
can make a valuable contribution to the avoidance of crack-
prone microstructures during in-service welding. The second
model presented in this paper demonstrates the accuracy with 2. Experimental methods
which hardness values can be predicted in simulated HAZ
structures, which in turn can be used to avoid crack-prone 2.1. Artificial neural network modelling
structures (>350 HV). This neural network model incorpo-
rates materials characteristics (chemical composition and Neuralworks Professional II/Plus Software was used to
as-received hardness), peak temperature, holding time and build the three back-propagation neural networks. As men-
cooling rate as key input variables for predicting hardness. tioned, the models were designed to predict the: (i) impact
The third model predicts hot ductility and hot strength toughness of quenched and tempered steels exposed to
of microalloyed steels based on chemical composition and repeated PWHT cycles (Model 1), (ii) simulated HAZ hard-
treatment conditions. The purpose of this model is to assess ness of pipeline and tap fitting steels after in-service welding
the likelihood of steels developing transverse cracking during (Model 2) and (iii) reduction of area (hot ductility) and hot

Table 1
Input variable data ranges used for the three models
Input variable Model 1 Model 2 Model 3

Minimum value Maximum value Minimum value Maximum value Minimum value Maximum value
Thickness (mm) 11 20
Test typea T L M S
Test temperature −90 20 700 1100
Timeb 0 200 0 2
Coolingc Slow Fast 2 10 100 200
Normalised HV 159 234
%C 0.155 0.180 0.07 0.29 0.155 0.165
%Mn 1.10 1.43 0.44 1.63 0.63 1.23
%S 0.0020 0.0035 0.005 0.024
%Cr 0.016 0.2 0.012 0.31
%B 0.0005 0.0013
%Si 0.005 0.39
%Cu 0.010 0.28
%Ni 0.009 0.1
%Mo 0.005 0.033
%V 0.005 0.06
%P 0.008 0.02
%Ti 0.005 0.021 0.003 0.018
%Al 0.005 0.039 0.027 0.036
%Nb 0.005 0.053 0.001 0.022
%N 0.0021 0.0033
a ‘T’ denotes the direction of rolling is transverse to the length of Charpy sample, ‘L’ denotes the direction of rolling is longitudinal to the length of Charpy

sample, ‘M’ denotes the samples were melted and solidified in situ to simulate direct casting conditions and ‘S’ denotes the samples were solution treated to
simulate re-heated conventional type cast structures.
b In Model 1 ‘time’ is PWHT time (min) and in Model 2 time is holding time (s) at 1400 ◦ C.
c In Model 1 ‘slow’ is cooling from 570 ◦ C at 250 ◦ C/h to 400 ◦ C and then still air, ‘fast’ cooling is cooling in still air from 570 ◦ C. In Model 2 ‘cooling’ is

t8/5 cooling time (s) and in Model 3 cooling is the cooling rate (◦ C/min) of the samples.
538 Z. Sterjovski et al. / Journal of Materials Processing Technology 170 (2005) 536–544

Table 2
Key neural network model parameters for the three back-propagation models
Model 1 Model 2 Model 3
Hidden layers 1 1 1
PEs in hidden layer 5 12 4
PEs in input layer 10 16 9
PEs in output layer 1 1 2
Learn rule Delta rule Delta rule Delta rule
Fig. 1. Schematic representation of a back-propagating ANN. Six and three Learn sequence Random Random Random
processing elements are shown, respectively, in the input and hidden layers. Learn cycles 150000 500000 500000
Transfer mode Sigmoid Hyperbolic tangent Sigmoid
strength of microalloyed steels during straightening in the Cases in training file 73 63 177
Cases in testing file 13 12 44
continuous casting process (Model 3). The input variables Epoch size 73 63 170
and ranges used for the models are shown in Table 1. For Learning coefficient (HL) 0.80 0.80 0.75
all models the input variables can be classified into two Learning coefficient (OL) 0.40 0.60 0.45
major areas: original material characteristics (chemical com- Learning coefficient ratio 0.85 0.85 0.80
position, plate thickness and/or hardness) and process/test Transition point 10000 10000 10000
Momentum 0.35 0.35 0.35
conditions (temperature, cooling rate and/or holding time).
Three data files were used in the modelling process,
namely the training file, testing file and the sensitivity file. The reached. The network type back-propagation derives its name
data are normalised between 0 and 1, or −1 and 1 depending from this method of calculating the error term [7].
on the intended transfer function, which modifies the sum of

the weightings into an output. For sigmoid transfer the data  
1
are normalised between 0 and 1, and for hyperbolic tangent RMSi =  (P(ij) − Tj )2 (1)
transfer the data are normalised between −1 and 1. The sen- n
j=1
sitivity file is generated by the neural network software upon
request. where P(ij) is the value predicted by the individual program
The initial step in building the models was to set the num- i for sample case j (from n sample cases) and Tj is the target
ber of processing elements (Fig. 1). Individual processing value for sample case j [8].
elements in the input layer correspond to an input variable Random weightings are assigned to each processing ele-
(listed in Table 1), and the processing element in the out- ment as an arbitrary starting point in the training process
put layer corresponds to a desired output. The role of the and the weightings are progressively altered on exposure to
hidden layer, which is a layer that contains a systematically numerous repetitions of training examples. Hence, the neural
determined number of processing elements, is to establish network is in the process of learning and constant adjustment
the relationships and inter-relationships between processing of weightings between processing elements leads to a reduc-
elements in the input and hidden layers that lead to the best tion in the RMS error of the overall training data. In the three
estimates of output. The latter stages in building the model models, 150,000–500,000 cycles were used to minimize and
include defining other parameters, such as selecting the learn stabilize the RMS error. The model reaches its optimum when
rule (the mathematical method the program uses to corre- the RMS error has stabilized. The model is verified against
late the processing elements in the entire model), learning cases in the test data file, which are independent of cases in
coefficients (values assigned to the output and hidden layers, the train data file. The predicted results are plotted versus the
and also the overall model) and transfer mode (transferring experimental (or actual) results.
the weighted sum of the processing elements into an output Finally, sensitivity data were used to further validate the
prediction) (see Table 2). Table 2 also lists the number of model and to interrogate the model to try to gain a physical
training cases used for each model and the epoch size, which understanding of the predicted trends. The sensitivity factor
is the number of training cases the model goes through before (SF) measures the contribution of each input variable to the
it readjusts the weightings of the processing elements that desired output, and it differs for each individual case in both
determine the value of the output. the train and test data files. Therefore, the SF presented in this
The learn rule used in each model is the delta rule and paper is an average of the contribution of an individual input
the error in the output layer is the difference between the variable towards the specified output. The SF for each data
desired output and the actual output. The error calculated case is generated by the Neuralworks Professional II/Plus
for the model was the root mean square (RMS) (Eq. (1)). Software, and this is then averaged to give an indication of the
This error, transformed by the derivative of the transfer average SF of an input variable. For further insight, the SF for
function, is “back-propagated” to prior layers where it is temperature for each prediction of ROA in the training file is
accumulated. This back-propagated and transformed error plotted versus test temperature in Model 3, and the sensitivity
becomes the error term for that prior layer. The process of factors for carbon and niobium for each case in the training
back-propagating the errors continues until the input layer is file are plotted versus the SF for carbon in Model 2.
Z. Sterjovski et al. / Journal of Materials Processing Technology 170 (2005) 536–544 539

2.2. PWHT simulation and impact toughness of Conditions for hot tensile tests were chosen to simulate both
quenched and tempered pressure vessel steels for the conventional casting (200–250 mm thick slabs) and thin slab
data in Model 1 casting (50–100 mm thick slabs) conditions. A Gleeble 3500
thermomechanical simulator was used to perform laboratory
PWHT of 11 mm 700PV grade, 12 mm 700 grade and tensile tests, which have been shown to correlate well with
20 mm 700PV grade quenched and tempered steels was car- the problem of transverse cracking [10,11].
ried out in a box furnace. Heat treatment conformed to Four microalloyed steels (Nb, Nb–Ti, Ti and C–Mn–Al) of
AS4458-1997-Pressure Equipment Manufacture, and was various chemical compositions were examined and the ten-
conducted in an argon atmosphere. The temperature of sile test specimens were cylindrical with a 10 mm diameter
treatment was 570 ± 5 ◦ C (validated by two thermocouples and a gauge length of 95 mm. Some samples were melted in
attached to each plate), and the holding time was 30 min for situ, under a compressive strain and quartz tube containment
the 11 and 12 mm plates and 50 min for the 20 mm plate. The to maintain the cross-sectional area of the melted sample.
ramp-up rate was 200 ◦ C/h and samples were then cooled in Other samples were solution treated at 1330 ◦ C. In both cases,
argon at 250 ◦ C/h or in still air. cooling rates of either 100 or 200 ◦ C/min down to the test tem-
Charpy V notch samples were machined and tested to perature range of 1100–700 ◦ C, corresponding to typically
AS1544.2-1989 from the middle of the plate. Parent plate used straightening temperatures, were used. Samples were
impact test data, transverse and longitudinal to the direction held for 1 min at the test temperature and then strained to
of rolling, were collected to build an artificial neural network failure at approximately 7.5 × 10−4 s−1 . After failure, sam-
capable of predicting impact toughness as function of PWHT ples were immediately water quenched. A minimum of four
time, chemical composition and test conditions. Five samples measurements were made on each fracture half to determine
were used for each group of tests. the reduction of area. The ultimate tensile strength (MPa) is
referred to as the hot tensile strength in this paper.
2.3. HAZ simulation and hardness testing of in-service
welding of pipeline and tap fitting steels for the data in
Model 2 3. Results

The simulation of the HAZ was carried out in a quench 3.1. Model 1
dilatometer. Tubular dilatometer samples with 5 mm outside
diameter, 3.5 mm inside diameter and 10 mm in length were Fig. 2 shows a plot of predicted impact energy versus
held between two quartz tubes and heated by a water-cooled experimental impact energy for both the train and test data for
induction coil under vacuum. A Type ‘S’ thermocouple spot- Model 1. There is no overlap between the two sets of data and
welded to the sample was used to monitor and control the good predictions occur in both cases. The RMS error of the
temperature of the sample. Eleven samples of various grades predictions for the entire data (train and test data combined)
(up to X70 grade) of Australian and USA pipeline and tap is 6.23 J.
fitting steels were heated to a peak temperature of 1400 ◦ C and
cooled at various rates (t8/5 of 2–10 s) to simulate the coarse 3.2. Model 2
grained HAZ resulting from in-service welding. The t8/5 (s)
is a commonly used parameter in the pipeline industry and it Fig. 3 shows a plot of predicted HAZ hardness versus
is defined as the time taken for a weldment to cool from 800 experimental HAZ hardness for both the train and test data
and to 500 ◦ C during in-service welding, and the equivalent for Model 2. There is no overlap between the two sets of data
linear cooling rate was maintained down to 150 ◦ C. Although and excellent predictions occur in both cases. The RMS error
typical arc-welding process thermal cycles ramp-up rates are of the predictions for the complete data is 11.69 HV.
at least 200 ◦ C/s [9], the ramp-up rate for HAZ simulation
was 50 ◦ C/s due to the limitations of the dilatometer. Holding
times were varied (0–2 s) to observe the effect of holding time
on hardness. Finally, Vickers hardness testing was carried
out using loads of 5 kg (as-received material) or 10 kg (HAZ
simulated material). Five indentations were averaged for each
type of test.

2.4. Hot ductility and hot strength of microalloyed steels


for the data in Model 3

Transverse cracking in microalloyed steels during


straightening in the continuous casting process is a problem Fig. 2. Predicted impact energy vs. measured impact energy for Model 1
often encountered in industry, especially for thin slab casters. (RMS error of 6.23 J).
540 Z. Sterjovski et al. / Journal of Materials Processing Technology 170 (2005) 536–544

Fig. 3. Predicted HAZ hardness vs. measured HAZ hardness for Model 2
(RMS error of 11.69 HV).

3.3. Model 3

Figs. 4 and 5, respectively, show a plot of predicted ROA


versus experimental ROA, and a plot of predicted hot tensile
strength versus experimental hot tensile strength, for both the
train and test data for Model 3. There is no overlap between
the two sets of data and reasonable predictions occur in both
cases, particularly for hot tensile strength (Fig. 5). The RMS
errors for the predictions for the entire data set are 7.79% for
ROA and 8.68 MPa for hot tensile strength.

3.4. Sensitivity results

Fig. 6 is a plot of the average SF for selected key input


variables in each ANN model. Positive values of the SF in

Fig. 6. SF vs. key input variables in the train data for (a) Model 1, (b) Model
2 and (c) Model 3. The average SF values error bars indicating a scatter of
±1 S.D.

Fig. 6 indicate an increase in property value with an increas-


Fig. 4. Predicted ROA vs. measured ROA for Model 3 (RMS error of 7.79%). ing value of the input variable. Negative values of SF predict
a decrease in property value with an increasing value of the
input variable. Fig. 7(a) is a plot of actual ROA versus test
temperature and Fig. 7(b) is a plot of the SF for test tem-
perature for each data case in the prediction of ROA and hot
strength of Model 3, which has a large standard deviation
for the prediction of ROA (see Fig. 6(c)) due to negative and
positive sensitivity factors. The calculated sensitivity value
varies for each data case and the scatter range is also shown
in Fig. 6 for selected variables from the three models. Fig. 8
is a plot of the SF in each data case for carbon versus the
SF for either original hardness or niobium in Model 2. This
graph illustrates a strong inter-relationship between the sen-
sitivity factors for carbon level and as-received hardness, and
Fig. 5. Predicted hot tensile strength vs. measured hot tensile strength for a non-definitive relationship between the factors for niobium
Model 3 (RMS error of 8.68 MPa). and carbon.
Z. Sterjovski et al. / Journal of Materials Processing Technology 170 (2005) 536–544 541

Fig. 9. Comparison of the predicted and measured HAZ hardness values for
the ANN model and the regression equation proposed by Yurioka et al. [14].

all error of the entire Charpy data, which was 7.5 J (standard
deviation) [12].

Fig. 7. (a) Typical ROA (%) vs. test temperature and (b) test temperature SF 4.2. Model 2
for ROA and hot strength (obtained from train data) vs. actual test tempera-
ture (Model 3).
A back-propagation ANN using the delta learn rule and
hyperbolic tangent transfer function was used to accurately
predict the simulated HAZ hardness values of a number of
pipeline and tap fitting steels with respect to 16 input param-
eters. The input parameters for this model included various
alloying elements, cooling time (t8/5 ) and holding time at
peak temperature (see Table 1). The levels of agreement for
the training data and test data sets are shown in Fig. 3. The
RMS error for the predictions was quite low, 11.69 HV, even
more apparent when the results of this model are compared
to results of other models published in the literature [13,14].
For example, a regression equation developed by Yurioka
et al. [14] is relatively poor in predicting the current data
Fig. 8. Plot of sensitivity factors of carbon (C), as-received hardness (HV)
(Fig. 9). This equation is based on Tekken y-groove welding
and niobium (Nb) in relation to predicted hardness for the train data of
Model 2. The SF for carbon, SF(C), is plotted against the sensitivity factors tests and therefore different cooling conditions to those asso-
for hardness, SF(HV), and niobium, SF(Nb). ciated with the dilatometer data. Consequently, the predicted
hardnesses are systematically lower. However, the scatter in
predicted hardness is much higher than that of the neural
4. Discussion network model.

4.1. Model 1 4.3. Model 3

ANN models have, to varying degrees, been successfully In the last case, a back-propagation neural network model
used to predict various properties of steels in relation to mate- using the delta learn rule and sigmoid transfer function was
rials characteristics and process/test conditions. In Model 1, successfully used to predict two outputs, ROA (hot ductility)
a back-propagation neural network using the delta learn rule and hot tensile strength, with respect to a number of input
and sigmoid transfer function has been employed to success- parameters as listed in Table 1. The levels of agreement for
fully predict impact energy in relation to a number of input the training data and test data sets are shown in Figs. 4 and 5.
parameters. The input parameters for this model included The RMS error for the prediction of ROA is significantly
material thickness, cumulative PWHT time, PWHT cooling higher than that for hot tensile strength, with the RMS error
type, C content, Mn content, S content, Cr content and B for the complete data set at 7.79% for ROA (15.6% error
content, test orientation and test temperature (Table 1). The on average ROA of 50%) and 8.68 MPa for the hot tensile
levels of agreement for the training data and test data sets strength (8.7% error on an average strength of 100 MPa). The
are shown in Fig. 2. The errors in predictions for both sets of scattered appearance in the ROA results in Fig. 4 is predom-
data are relatively low, with the RMS error for the combined inantly due to the method used to measure reduction of area
data set at 6.23 J. This error is lower than the average over- (vernier calliper measurements of diameter) and the reduced
542 Z. Sterjovski et al. / Journal of Materials Processing Technology 170 (2005) 536–544

confidence in the accuracy of the melt samples due to the tures undergoing autotempering during cooling through the
difficult nature of the test. Ms-Mf range, or the autotempering of bainitic structures
formed at higher temperatures. At still lower cooling rates
4.4. Sensitivity analysis softer transformation products will be produced (such as fer-
rite and bainite formed at higher temperatures).
Sensitivity analysis was used to further validate each Test temperature has the most significant effect on impact
model and examine the contribution of an input variable to the energy of all the input variables in Model 1 (Fig. 6(a)). In
output for each individual case in the train data. Fig. 6 shows its learning phase, the model was able to predict a ductile to
the average SF of selected input variables for each model. brittle transition temperature curve [12], hence further val-
As mentioned, positive values of the SF in Fig. 6 indicate an idating the model. A decrease in hot tensile strength with
increase in the value of the output with an increasing value an increase in test temperature is also an expected result in
of the input variable. Negative values of the SF predict a Model 3 (Fig. 6(c)), due to increased mobility of dislocations
decrease in the value of the output with an increasing value at higher temperatures associated with decreased resistance
of the input variable. to dislocation glide and climb, and the generation of a failure
Fig. 6 shows that carbon content is a significant input vari- mechanism at a lower stress level [21].
able for all of the models. Carbon is known to have a strong Test temperature has a complex effect on the ROA, evi-
influence on hardness, strength and ductility [15]. Increasing dent by the large value for standard deviation in the SF for
the carbon content of steel leads to increased hardness and test temperature (Fig. 6(c)). This large scatter is due to the
reduced ductility [16], and hence reduced impact energy. The presence of different phases at temperatures below ∼900 ◦ C
carbon content in neural network Model 1 ranged from 0.15 compared to temperatures above ∼900 ◦ C. Below ∼900 ◦ C,
to 0.19%, and the average sensitivity calculated across this an increase in test temperature in the presence of dual phases
range was −8.6% (Fig. 6(a)), which implies that an increase (ferrite and austenite) results in a decrease in ROA (indicated
in C content decreases the impact energy (over the specified by the predominantly negative SF values in Fig. 7(b)). As test
C range). temperature moves away from 900 ◦ C and towards 700 ◦ C
The carbon content values used in Model 2 ranged from the sensitivity factors in Fig. 7(b) indicate an increased neg-
0.07 to 0.29 wt%, and the SF calculated along this range was ative effect on ROA. At temperatures where austenite is the
85 (Fig. 6(b)). This predicted effect of carbon on hardness in only phase present, above ∼900 ◦ C, the sensitivity factors on
the model is confirmed by Irvine and Pickering [17]. Their average show that an increase in test temperature results in
work shows a significant increase in hardness with increas- an increase in ROA (indicated by the positive SF values in
ing carbon content of various steels that have martensitic and Fig. 7(b)). This trend is well documented and is schemati-
bainitic microstructures, which are the dominant microstruc- cally shown in Fig. 10 [10]. The high sensitivity of SF to test
tures of the simulated HAZ samples in the current work. temperatures between 900 and 700 ◦ C reflects data associated
Increasing carbon content increases the hardenability of steel with variable volume fractions of ferrite and austenite, and
[18], thus reducing the capacity to form softer transformation indicates that caution should be exercised in predicting trends
products (such as upper bainite and ferrite) and promoting the in ROA from the model over this range of temperatures.
formation of a predominantly martensitic structure, which An in-depth understanding of ANNs is required to use
is significantly harder than ferrite or upper bainite. In fully them effectively as design tools and a comprehensive under-
martensitic structures, increases in hardness with increasing standing of the subject being modelled is also required.
carbon content are mainly due to the increased concentration Analysis of the neural network model can become difficult
of carbon in supersaturated solution [19]. Carbon is inter-
related with the hardness of the as-received (normalised)
steels (Fig. 8) and the consequences of this relationship are
discussed later.
Temperature and holding time at temperature typically
have a significant impact on the mechanical properties of
steels. Fig. 6(a) shows that impact energy decreases with
cumulative PWHT time, and this is attributed to significant
coarsening of second phase carbide particles, which accel-
erates high strain rate fracture [20]. In Model 2, increasing
cooling time, t8/5 (s), from 2 to 10 s (or decreasing cooling
rate from 150 to 30 ◦ C/s) results in a SF of −17 (Fig. 6(b)).
This suggests that slowing down the cooling rate (increasing
cooling time) will result in a significant drop in hardness,
although the effect is not as significant as decreasing carbon
content. The reduction in hardness as a result of decreasing Fig. 10. Schematic diagram showing the three characteristic ductility
cooling rate is due predominantly to martensitic microstruc- regions on a hot ductility curve for a low carbon steel [10].
Z. Sterjovski et al. / Journal of Materials Processing Technology 170 (2005) 536–544 543

in the presence of inter-related input variables. Optimizing service welding. Finally, Model 3, which is based on data
parameters (process or design) when input variables are not generated from experimental simulations, can assist in deter-
inter-related is simply a matter of determining the SF for a mining the optimal temperature to avoid transverse cracking
particular variable and appropriately reducing or increasing in microalloyed steels during straightening. The potential
the magnitude of this variable to achieve the desired out- for transverse cracking, in this situation, can be minimized
put. However, interdependent input variables can introduce a by maximizing hot ductility (increasing ROA) and optimiz-
higher degree of complexity in the interpretation of sensitiv- ing the temperature of straightening in the ANN, which is
ity data. also effected by the composition of the microalloyed steel.
Fig. 8 shows the SF for wt% carbon associated with From a practical viewpoint, ANNs can be used to predict the
the train data of Model 2 plotted against the SF for as- likelihood of transverse cracking in the actual straightening
received steel hardness, and also against the SF for niobium. process. For example, a characterization model could be used
Although the variables of C and Nb contents are independent, to predict the likelihood of transverse cracking based on the
an inter-relationship exists between carbon and as-received input variables used for Model 3 (materials composition, test
steel hardness. As-received hardness is a property that relates temperature and cooling rate) and other casting variables,
quantitatively to composition, microstructure, grain size and such as slab thickness, mould exit temperature and casting
segregation effects. In particular, the hardness of these nor- speed.
malised low alloy steels will largely depend on carbon content
(and hence volume fraction of pearlite). The presence and
absence of inter-relationships between input variables are 5. Conclusions
indicated by the relationships between the sensitivity fac-
tors in Fig. 8: direct proportionality for SF(C) and SF(HV), For the ranges of the input variables examined the three
essentially no relationship for SF(C) and SF(Nb). Although back-propagation neural network models successfully pre-
the incorporation of both as-received hardness and C content dicted impact toughness, HAZ hardness, ROA and hot tensile
as input variables might appear to be redundant and unnec- strength for variations in as-received material characteristics
essary, the as-received hardness was incorporated into the and process/test parameters. It was also shown that ANNs
model to provide a more complete set of input variables that could successfully predict multiple mechanical properties.
relate to the starting material characteristics and hence are rel- The results of the sensitivity analysis were in agreement
evant to the accurate prediction of HAZ hardness. Therefore, with both findings of the experimental investigations and
the quantitative contribution of carbon to HAZ hardness was reports in the literature. Manipulation of the sensitivity data
considered by using both the carbon content and the start- showed the complex nature of ANNs and their ability to solve
ing (normalised) hardness as input variables, with the aim complex patterns. The presence of inter-related variables does
of gaining improved insight into the carbon levels required not compromise model predictions, but interpretation of the
to achieve an acceptable HAZ hardness value. This exer- predicted trends in sensitivity data must be approached with
cise demonstrates that the interdependence of input variables caution to effectively optimize design or process parameters.
does not compromise the capacity of the ANN model to pre- ANNs have the potential to minimize the need for expen-
dict HAZ hardness, nor to establish, for the purpose of steel sive experimental investigation and/or inspection of steels
design, the carbon level required to limit the HAZ hardness used in various applications, hence resulting in large eco-
level, for a given cooling rate. nomic benefits for organisations.
In summary, ANNs can eliminate the need for detailed
experimental investigation into processes and even other
Acknowledgements
areas, such as inspection and field trials, through their abil-
ity to accurately predict the mechanical properties of steels.
The current work was conducted as part of projects spon-
Model 1 shows that if the history of relevant data for in-
sored by the Cooperative Research Centre for Welded Struc-
service pressure vessels (PWHT temperature, cumulative
tures (CRC-WS) and its core partner, the Australian Pipeline
PWHT time, chemical composition, wall thickness, etc.)
Industry Association (APIA). CRC-WS was established and
is recorded, then the need for inspection and testing may
is supported by the Australian Government Cooperative
only be required if predicted impact toughness values reach
Research Centres Program. The authors would also like to
or approach non-conformance limits in the relevant Aus-
thank the Australian Research Council and Bluescope Steel
tralian standards. Similarly, Model 2 shows that through
for their support.
prior knowledge of material composition and the relation-
ship between weld heat input and weldment cooling rates
(also dependent on pipe wall thickness and internal fluid flow
References
rates), neural networks can predict HAZ hardness to provide
knowledge of susceptibility of microstructure to cracking. [1] D. Graupe, Principles of Artificial Neural Networks—Advanced
This can therefore eliminate the need for extensive in-service Series on Circuits and Systems, vol. 3, World Scientific Publish-
welding trials and/or hardness testing of each weld after in- ing, Singapore, 1997.
544 Z. Sterjovski et al. / Journal of Materials Processing Technology 170 (2005) 536–544

[2] S. Alter, Information Systems—A Management Perspective, third ceedings: Thermomechanical Processing: Microstructure, Modelling
ed., Addison Wesley Longman, 1999. and Control, IMMPETUS, University of Sheffield, UK, June 23–26,
[3] T. Cool, H.K.D.H. Bhadeshia, D.J.C. MacKay, The use of an arti- 2002, pp. 270–274.
ficial neural network to investigate the yield and ultimate tensile [12] Z. Sterjovski, D. Dunne, S. Ambrose, Modelling of repeatedly post-
strength of steel welds, Mater. Sci. Eng. A A223 (1–2) (1997) weld heat-treated quenched and tempered steel, Australas. Weld. J.
186–200. 48 (2nd Quarter) (2003) 41–46.
[4] L.X. Kong, P.D. Hodgson, D.C. Collinson, ISIJ Int. 38 (10) (1998) [13] M. Painter, In-Service Welding of Gas Pipelines—Project 96:34 Final
1121–1129. Report, Cooperative Research Centre for Welded Structures, Wollon-
[5] H. Tsuei, Analysis and Modelling of Weld Metal Mechanical Prop- gong, Australia, June 2000.
erties in Flux Cored Arc Welded Steels, PhD Thesis, University of [14] N. Yurioka, H. Suzuki, S. Oshita, Weld. J. 62 (6) (1983) 147–153.
Wollongong, 2000. [15] American Society for Metals, Metals Handbook, vol. 6, ninth ed.,
[6] Z. Sterjovski, D. Nolan, D. Dunne, J. Norrish, Predicting the HAZ 1983.
hardness of pipeline and tap fitting steels with artificial neural net- [16] C.A. Siebert, D.V. Doane, D.H. Breen, The Hardenability of
works, in: Proceedings of the Fourth International Conference on Steels: Concepts, Metallurgical Influences and Industrial Applica-
Pipeline Technology, vol. 3, Ostend, Belgium, May 9–13, 2004, pp. tions, American Society for Metals, Metals Park, OH, 1977.
1223–1233. [17] K.J. Irvine, F.B. Pickering, J. Iron Steel Inst. 187 (1957) 292.
[7] Neural Ware Incorporated, Neural Computing Guide, Technical Pub- [18] R.W.K. Honeycombe, Steels: Microstructures and Properties, Edward
lications Group, Pittsburgh, U.S.A., 2003. Arnold, London, 1981, p. 131.
[8] http://www.gepsoft.com/gepsoft/APS3KB/Chapter09/Section1/SS04. [19] D.R. Askeland, The Science and Engineering of Materials, second
htm, November 2004. ed., Chapman and Hall, London, 1990, p. 263.
[9] J.F. Lancaster, Metallurgy of Welding, sixth ed., Abington Publish- [20] Z. Sterjovski, D. Dunne, S. Ambrose, Multiple PWHT of submerged
ing, Cambridge, 1999. arc welded quenched and tempered pressure vessel steels, in: Con-
[10] B. Mintz, S. Yue, J.J. Jonas, Hot ductility of steels and its relationship ference Proceedings: International Institute of Welding (IIW) Asian
to the problem of transverse cracking during continuous casting, Int. Pacific International Congress, Suntec Centre, Singapore, October
Mater. Rev. 36 (5) (1991) 187–217. 2002.
[11] K.R. Carpenter, B.A. Parker, C.R. Killmore, Hot ductility of Nb, [21] R.E. Reed-Hill, R. Abbashian, Physical Metallurgy Principles, third
Ti and Nb–Ti microalloyed C–Mn–Al steels, in: Conference Pro- ed., PWS-KENT, Boston, 1992.

You might also like