You are on page 1of 7

Materials Technology

Investigation of Decarburization in Spring Steel Production Process – Part II: Simulation


Dejun Li1,2), D. Anghelina3), D. Burzic1), W. Krieger3), E. Kozeschnik4)

1)
Materials Center Leoben Forschung GmbH, Roseggerstrase 12, 8700 Leoben, Austria; denijel.burzic@tuwien.ac.at
2)
now at Institute of Mechatronic Control Engineering, College of Mechanical and Energy Engineering, Zhejiang University, 38 Zheda Road,
Hangzhou, ZheJiang, 310027, P.R. China
3)
Chair of Metallurgy, University of Leoben, Franz-josef-Str. 18, 8700 Leoben, Austria.
4)
Institute of Materials Science and Technology, Vienna University of Technology, Favoritenstr. 9-11/E308, 1040 Wien, Austria

In part I of this paper, the decarburization and oxidation behavior of spring steel during simulated thermal cycles resembling the
conventional production process have been investigated experimentally. In part II, the results obtained from part I are studied theoretically
and numerically by various computational methods. The phenomena discussed in this study include the influence of composition on phase
transformation and on diffusion behavior of carbon, decarburization process calculations by simulations of diffusion controlled phase
transformations with the software DICTRA, and the development of a simple integrated model to describe simultaneously the
decarburization and oxidation kinetics. The simulations show good agreement with experimental results. Moreover, the simulation
methodologies can be used to optimize processing parameters and steel composition.

Keywords: decarburization, oxidation, spring steel, simulation

DOI: 10.2374/SRI08SP070; submitted on 18 March 2008, accepted on 29 October 2008

Introduction Decoupled Treatment of Oxidation and Decarburization

Surface decarburization in the process of spring steel Investigation of surface oxidation


production often leads to deteriorated mechanical behavior
of the finished components in service. The metallurgical The kinetics of oxidation and the stability of the
processes are complex and influenced by many factors, oxidation products are strongly dependent on the type of
such as temperature, atmosphere, D/J transformation, oxide forming element and, consequently, of the steel
morphology of the surface, alloy composition and many composition. Fayalite is a typical silicon-rich oxide in Si-
more. These phenomena have been investigated many containing steels and it easily forms with a strong adhesion
times [1-3]. In this study, decarburization behavior of the between oxide scale and substrate, which can lead to
conventional Si-rich spring steel 54SiCr6, has been increased resistance against oxidation. On the other hand,
examined, theoretically and numerically based on the Si and its compounds are relatively rigid materials and
experimental data obtained on the Gleeble 1500 thermo- may act as stress concentration sites. This may lead to
mechanical simulator in part I [22] of this paper. cracking of the scale. Although it was reported that the
To understand the mutual interaction between oxidation second effect seems to be stronger, it is important to
and decarburization, computer simulation can be very acquire better quantitative knowledge about the influence
helpful and assist in interpretation of complex and of steel composition on oxide formation [13-15].
sometimes counter-intuitive results. Simulation is often Figure 1 shows the equilibrium phase fraction diagram
cheaper and faster than experimental investigation, and in of 54SiCr6 as a function of oxygen activity calculated with
many cases, can provide not only trends but also ThermoCalc and the TCFE3 database [12]. This type of
quantitative information on the material behavior under plot provides information on the type of oxide which
laboratory as well as industrial condition. In this context, forms with increasing oxygen content from the steel
several models have been developed to investigate the substrate (left, low O activity) through the entire oxide
decarburization process [4-8]. layer to the gas atmosphere (right, highest O activity). For
In the present work, computer simulations have been the given steel composition and temperatures exceeding
used to investigate the influence of the chemical 1100°C, it is evident that a considerable amount of fayalite
composition of the steel on the formation of oxide layers forms (approximately 10 mass-%). The majority of all
and decarburization in steels based on phase oxides are present in the form of iron oxides, which
transformation and diffusion calculation, and on the change with increasing O activity from FeO to Fe3O4 and,
decarburization phenomenon during typical industrial finally, Fe2O3 on the scale/gas interface.
conditions based on the finite differences approach. According to the thermodynamic equilibrium calcula-
Different software has been used in this purpose such as tions, most of the fayalite phase is formed within the FeO
MatCalc [9-11], ThermoCalc and DICTRA [12], but also a layer. With increasing oxygen activity, Fe3O4 starts to
special program designed to simulate the decarburization form. A considerable amount of fayalite can also be
and oxidation phenomena simultaneously. This software observed within the Fe3O4 layer, but it can only be found
has been utilized to obtain a more comprehensive picture in the regions next to the FeO layer. With higher oxygen
of the interaction of the individual metallurgical processes activity, fayalite is replaced by SiO2, which is stable both
in laboratory and industrial environment. in the Fe3O4 and Fe2O3 layers.

304 steel research int. 80 (2009) No. 4


Materials Technology

Figure 1. Oxide weight ratio in equilibrium in steel 54SiCr6 (0.55C- Figure 2. Oxide weight ratio in equilibrium in steel 51CrV4 (0.51C-
1.5Si-0.65Cr-0.65Mn, in mass-%)) as function of oxygen activity. 0.3Si-1.0Cr-1.0Mn-0.2V, in mass-%) as function of oxygen activity.

Figure 3. Influence of Cr, Mn and Si on A1 and A3 temperature of steel 54SiCr6 (other elements kept at nominal composition).

For comparison, the phase fraction diagram of another this section, the influence of the major alloying elements
steel 51CrV4 is shown in Figure 2. This steel contains on the equilibrium phase transformation temperatures and
only small amounts of silicon and it is apparent that, due to the carbon diffusion coefficient are investigated using the
the low concentration of silicon, only a small quantity of software and databases of MatCalc [9-11] version 5.21.
fayalite forms with a mass percentage of less than 0.05%.
However, more Cr2O3 and Mn2SiO4 form due to the higher Changes in Ae1 and Ae3 temperatures for Ȗ ĺ Į trans-
concentration of Cr and Si in this steel compared with formation. In Figure 3 the influence of the Si, Cr and Mn
54SiCr6. These computational results confirm the content on the Ae1 and Ae3 temperatures of the steel
experimental evidence that different chemical composition 54SiCr6 is summarized. Accordingly, Cr increases Ae1 but
leads to pronounced differences in the kind of oxides that decreases Ae3, Mn decreases both Ae1 and Ae3. Si increases
form. However, this investigation also indicates that there both Ae1 and Ae3. These effects are well known, as well as
is only little difference in the sequence of wustite (FeO), the fact that, at the same temperature, the diffusion
magnetite (Fe3O4) and hematite (Fe2O3) formation with coefficient of carbon in bcc iron is much higher than in fcc
activity of oxygen for steels 54SiCr6 and 51CrV4. iron. For the decarburization behavior of steel during
continuous cooling, this means that, if Ae1 and Ae3 are
Influence of elements on decarburization kinetics shifted to higher temperature, the temperature range of fast
diffusion in bcc is wider and more decarburization is
It is well known that alloying elements can strongly expected to occur during processing. From this viewpoint,
influence the steel decarburization behavior [16-18]. In it is obvious that Si can be expected to be detrimental if

steel research int. 80 (2009) No. 4 305


Materials Technology

operating mechanisms and conclusions from individual


processes to the overall material behavior must always be
considered with care.

The influence of elements on carbon diffusion. The


kinetics of decarburization is directly proportional to the
diffusional mobility of carbon in steel: A low carbon
diffusion coefficient leads to low amount of decarburize-
tion. Consequently, the influence of various elements on
the carbon diffusion coefficient is investigated next.
Figure 4 presents the influence of major alloying
elements Cr, Si and Mn on the carbon diffusion coefficient
in the austenite phase field of steel 54SiCr6 at 1000°C. It
is apparent that only high fractions of Cr, Mn and Si can
Figure 4. Influence of elements Cr, Si and Mn on the carbon decrease the diffusion coefficient of carbon in the steel
diffusion coefficient in steel 54SiCr6 (other elements kept at sufficiently to cause major effects. Among all elements, Cr
nominal composition).
produces the largest effect.

Numerical simulation of decarburization

The evolution of the carbon diffusion profile at the


contact area between metal and atmosphere under typical
production conditions (with some simplifications) for the
54SiCr6 spring steel have been simulated using the
commercial software DICTRA.

Simulation of the influence of Į ൺ Ȗ phase trans-


formation on decarburization. Figure 5 shows the
decarburized depth for a bcc and an fcc substrate at
different temperatures and after different times calculated
with DICTRA. As expected, the decarburized depth
increases with temperature for both D and J substrates and
decarburization proceeds much faster in bcc than in fcc. At
the same temperature and after the same time, the
Figure 5. Comparison of the decarburized depth between bcc decarburized depth in bcc iron is more than 50 times
substrate and fcc substrate (0.9 C0 as criterion). higher than in fcc iron.

Decarburization simulation for isothermal condition of


Experiment II (from part I). In this section, the DICTRA
software is applied to simulate the decarburization process
during our previous experiment II in Part I of this paper
(figure of applied thermo cycle is given in Part I). Since
the heating and cooling rate were very fast and martensite
formed during the cooling process in the experiment, it is
assumed that the decarburization process during heating
and cooling can be neglected and, consequently, have not
been included in our model. Therefore, only the isothermal
decarburization process for the specimens, which were
held at the high temperature of 1050°C, is simulated. The
substrate is assumed to be J phase during the entire
process.
Figure 6 presents the decarburized depth as a function
of annealing time. The simulations show that the
Figure 6. Comparison between experimental and simulation decarburized depth increases approximately parabolic with
results for Experiment II (Part I). increasing holding time.
When comparing with the experimental values obtained
in part I of this paper, it must be kept in mind that the
minimum decarburization is desired. However, it must be evaluation of the decarburized depth has been carried out
kept in mind that the simultaneous decarburization and by optically evaluating the microstructure images after
oxidation is determined by the combined effect of all quenching to room temperature. The decarburized depth

306 steel research int. 80 (2009) No. 4


Materials Technology

has been identified as the region where decarburization


becomes visible by an increase of the ferrite fraction, i.e.
the white-etching microstructure constituent. When
measuring the decarburized depth after 700 s with 100 µm,
it is easily found that this value corresponds to a position
with 0.2 wt% carbon concentration in our simulation. This
value corresponds to a value of 0.36 C0 (with C0 = 0.55
wt%). Therefore, in Figure 6, this reference value has been
compared to the DICTRA simulations and, apparently,
excellent agreement between simulation and experiment is
obtained.

Decarburization in the presence of the J/D phase trans-


formation. In the previous section, decarburization has
only been treated in the J one-phase region during
isothermal annealing at 105 °C. Now, the J/D phase
(a)
transformation is taken into account during continuous
cooling according to experiment III of part I of this paper.
The simulations are then compared to the experimental
results for specimens III-1 through III-4.
When setting up the DICTRA simulations, two different
approaches have been followed. In model A, it has been
assumed that during cooling only J phase exists. This
situation corresponds to steel with sluggish austenite
decomposition kinetics. In model B, it has been assumed
that D phase can form on the surface of the substrate when
temperature decreases or carbon concentration decreases
to some values which support the formation of D. The
precise moment when ferrite starts to form is calculated
automatically by DICTRA based on the Fe-C diagram and
the local equilibrium hypothesis across the interface.
The thermal cycles used in both models are identical to
Experiment III in Part I of this paper. Figure 7a and 7b
(b)
show simulation results for specimens III-1 though III-4,
calculated with model A and model B separately. Table 1 Figure 7. Simulation of the decarburization process in Experiment
shows the evaluated decarburization depth from III with influence of the J/D phase transformation: (a) Model A: No
Experiment III, as well as the predicted decarburization fcc/bcc phase transformation (b) Model B: Including fcc/bcc phase
depth from DICTRA simulations using the 0.36 C0 transformation.
criterion for model A and B.
If no D phase is allowed to form during cooling (model protecting the bulk metal from rapid loss of carbon by
A), there is no large difference in decarburized depth for sealing against the outer atmosphere.
specimens III-1 to 4. Most decarburization occurs during
the holding period at 1050°C. If the phase transformation
is included in the simulation (model B), after reaching the An Integrated Model for Simultaneous Oxidation and
A3 temperature, an D film forms on the surface with its Decarburization
typically low carbon content. According to the local
equilibrium model, carbon is continuously pushed ahead Under industrial conditions, oxidation occurs
the moving J/D phase boundary into the austenite phase simultaneously to the diffusion of carbon out of the metal
while the ferrite phase is growing. As soon as the carbon substrate. It is important to realize that, while these two
content in austenite reaches the eutectoid content of 0.8
wt-%, pearlite formation starts and the J/D phase
transformation comes to an end. Table 1. Comparison of decarburization depth between models A,
B and experimental results.
Model B shows reasonable agreement with the
experimental results. The values of the decarburization Decarburization depth [µm]
No.
depth are slightly larger for simulations compared to Experiment Model A Model B
experimental results. This is expected since the influence III-1 43 52.6 52.6
of oxidation during decarburization, and formation of III-2 51 62.2 62.2
oxide layer, is not taken into account. Oxide formation III-3 53 64.7 73.6
decreases the decarburized depth by transforming parts of III-4 85 65.0 107.0 (101)*)
the surface decarburized layer into oxide and, also, by *) calculation stopped at 0.8 wt% C-supersaturation in J

steel research int. 80 (2009) No. 4 307


Materials Technology

Modeling of oxide scale growth. In numerous experi-


ments [19,20] the growth of oxide layers on steel
substrates has been investigated. Depending on the type of
atmosphere and the amount of continuous oxygen support,
linear or parabolic growth laws have been observed. Since
fundamental and general growth relations for oxidation
processes in multi-component systems are lacking due to
the complexity of the subject and the numerous factors
influencing the process, in a recent publication [19], an
attempt has been undertaken to describe the growth of
oxide layers by either one of the following equations:

§W · (1)
¨ ¸ kl ˜ t
© A¹
Figure 8. Mass increase curves from oxidation of 54SiCr6 steel in 2
furnace gas at different temperatures. §W · (2)
¨ ¸ kp ˜t
© A¹

where W is the weight of the atomic oxygen [kg], A is the


surface area of the sample [m2], kl is the linear rate
constant [kg/m2·s], kp is the parabolic rate constant
[kg2/m4·s], and t is time [s].
Equations (1) and (2) describe linear and parabolic
growth rates, respectively. The initial stages of oxidation
usually obey the linear growth rate law, because oxide
growth is controlled either by the rate of chemical reaction
at the metal surface, or by the rate of adsorption of the
oxidizing species on the scale surface and the
incorporation of atomic oxygen into the forming layer.
When the oxide layer reaches a certain thickness, the
mechanism of oxidation becomes controlled by the
diffusion of ionic species through the oxide layer, leading
Figure 9. Mass increase comparison between calculated and to a parabolic growth rate [20].
experimental results.
In order to obtain accurate oxide scale growth rates,
isothermal oxidation experiments have been performed, in
concurring processes take place, the oxide scale growth the furnace gas and for the temperature range from 700 to
usually consumes part of the decarburized metal and thus 1200°C. Argon has been used during heating and cooling
decreases the finally observed decarburized depth. If the to provide inert conditions during the non-isothermal
decarburized depth is to be predicted, it is necessary to sections. Figure 8 shows the observed mass increase
include both processes in an integrated model and simulate curves, where the large dependence on temperature can be
scale growth and decarburization simultaneously. Since no noticed.
such computer procedure is readily available, a simple During calibration of the program, it has been found that
model has been developed, which is introduced in the fitting the mass increase curves with only one parabolic or
following sections. linear growth law sometimes produces large errors.
Therefore, curves which could not be described
satisfactorily with only one single linear or parabolic
Table 2. Values of the constants used in the integrated model. growth law have been fitted using two linear segments.
Table 2 shows values of the constants used in the
Temperature [°C] Values of constants [gs-1cm-2]
integrated model.
Kl1 Kl2
700 4.56·10-6 4.57·10-7 Figure 9 shows a comparison between the mass increase,
800 4.09·10 -6
6.70·10-7 observed experimentally and predicted from simulation.
900 6.74·10 -6
6.07·10-7 Good agreement is found between them, although slight
1000 2.05·10 -5
5.81·10-6 deviations can be noticed under some conditions. These
1050 2.75·10 -5
8.95·10-6 differences can be attributed to irregularities in the
1100 2.64·10 -5
1.07·10-5 oxidation process occurring during the experiments, which
1150 2.54·10 -5
-**) are difficult to include in the simulation model. Typical
1200 3.67·10-5 -**) examples of such processes are crack formation or partial
**) Experiments at 1150 and 1200°C described with one linear segment removal of the oxide layer.

308 steel research int. 80 (2009) No. 4


Materials Technology

Decarburized depth growth. In a rigorous approach,


the transient diffusion equation for one-dimensional semi-
infinite diffusion geometry can be solved for non-
isothermal conditions and multi-component environment
using, e.g., the software DICTRA. However, in our
simplified integrated model, a much more simple but
efficient approach has been chosen. Based on the vacancy-
exchange mechanism and random-walk theory, the
spreading of a diffusion cloud (SDC) can be approximated
by the mean square diffusion distance <r2> as [21]

r2 2dDt (3)

where r is the average radius of the spreading diffusion


cloud, d denotes the spatial dimensionality (d=1, 2 or 3), D
Figure 10. Decarburized depth and oxide depth at typical thermal
is the diffusion coefficient, and t is time. In the integrated
cycles in furnace gas.
model, the carbon concentration profile developing in the
course of the decarburization process is not explicitly
taken into account. Only the average distance that the
carbon atoms travel within a given time span, at a certain steels. An integrated model has been developed which
temperature and depending on the type of matrix (fcc or includes effects of both phenomena. The oxidation rate has
bcc) is considered. Under non-isothermal conditions, the been described using linear or parabolic growth laws (Eq.
total diffusion distance is straightforwardly obtained by (1) and (2)), while the decarburization depth has been
piecewise linear integration. The final decarburized depth estimated using an approximation for the random walk
is obtained as the difference between the total mean mean diffusion distance (Eq. (3)). The model shows good
diffusion distance and the loss of metal substrate by oxide agreement with experimental results, and can be used to
growth. The latter is calculated from the total oxide scale optimize the thermal cycle during steel production.
thickness and the mass balance taking into account the
type of oxide.
Figure 10 shows the decarburized depth for a typical Acknowledgement
thermal cycle during production of 54SiCr6 steel. It can be
noticed that the decarburized depth grows continuously Financial support by the Österreichische Forschungs-
with time, where increasing temperature leads to förderungsgesellschaft mbH, the Province of Styria, the
increasing growth rates. The decarburization rate is Steirische Wirtschaftsförderungsgesellschaft mbH and the
strongly influenced by the carbon diffusion coefficient. Municipality of Leoben within research activities of the
When the temperature reaches approximately 800°C, and Materials Center Leoben Forschung GmbH in the
the Jto D phase transformation starts, it can be noticed that framework of the Austrian Kplus Competence Center
the decarburization rate decreases. The explanation of this Programme is gratefully acknowledged.
phenomenon is that the carbon diffusion coefficient in
ferrite is much higher than that in austenite.
References

Conclusions [1] J. Gegner: Z. Metallkd., 94, (2003), 30-35.


[2] J. Kucera, P. Braz and K. Adamaszek: Acta Techn., 45 (2000), 45-64.
The results obtained from part I of this paper have been [3] J. Kucera and K. Adamaszek: Acta Techn., 46 (2001), 51-60.
[4] A. Phillion, H. S. Zurob, C. R. Hutchinson, H. Guo, D. V. Malakhov,
studied theoretically and numerically using several
J. Nakano, G. R. Prudy: Metall. Trans. A, 35A (2004), 1237-1241.
computational methods. Numerical simulations have in all [5] B. Millon, J. Ruzickova and K. Stransky: Kovove Mater., 27 (1989),
cases proved to be useful in determining trends in No.4, 492-503.
materials behavior during thermal treatment. Moreover, [6] C. R. Oldani: Scripta Mater., 35 (1996), No.11, 1253-1257.
quantitative information on the oxidation and decarburiza- [7] P. Marini and G. Abbruzzese: J. Magn. Magn. Mat., 26 (1982), 15-21.
[8] Y. Prawoto, N. Sato, I. Otanl and M. Ikeda: J. Mater. Eng.
tion characteristics has been obtained in good agreement
Performance, 13 (2004), No.5, 627-636.
with the experiments. [9] E. Kozeschnik, J. Svoboda, P. Fratzl and F. D. Fischer: Mater. Sci.
Simulations have shown that the chemical composition Eng., A 385 (2004), No.1-2, 157-165.
has a large affect on oxidation, as well as on decarburiza- [10] J. Svoboda, F. D. Fischer, P. Fratzl and E. Kozeschnik: Mater. Sci.
tion. The type of oxide that forms is strongly dependent on Eng., A 385 (2004), No.1-2, 166-174.
the steel chemical composition, particularly with regard to [11] E. Kozeschnik, J. Svoboda and F. D. Fischer: CALPHAD, 28 (2004),
No. 4, 379-382.
the effect of silicon. [12] J. O. Andersson, T. Helander, L. Höglund. P. Shi and B. Sundman:
Oxidation and surface decarburization occur Calphad, 26 (2002), No.2, 273-312.
simultaneously during the production process of spring [13] P. J. Szabo and E. Denes: Mikrochim. Acta, 59 (2003), 433-468.

steel research int. 80 (2009) No. 4 309


News

[14] S. Taniguchi, K. Yamamoto, D. Megumi and T. Shibata: Mater. Sci [19] H. T. Abuluwefa, R. I. L. Guthrie and F. Ajersch: Metall. Mater.
Eng., A 308 (2001), 250-257. Trans., A 28 (1997), 1633-1641.
[15] H. Okada, T. Fukagawa, H. Ishihara, A. Okamoto, M. Azuma and Y. [20] H. T. Abuluwefa, R. I. L. Guthrie and F. Ajersch: Oxid. Met., 46
Matsuda: ISIJ Int., 35 (1995), No.7, 886-891. (1996), No. 5, 423-440.
[16] A. Mayer: Eisen, 83 (1963), No. 19, 1169-1176. [21] M. E. Glicksman: Diffusion in Solids, John Wiley & Sons, New York,
[17] J. S. Dunning, D. E. Alman and J. C. Rawers: Oxid. Met., 57 (2002), 2000.
No.5, 409-425. [22] D. Li, D. Anghelina, D. Burzic, J. Zamberger, R. Kienreich, H.
[18] D. Geneve, M. Confente, D. Rouxel, P. Pigeat and B. Weber: Oxid. Schifferl, W. Krieger, E. Kozeschnik: Steel Research Int., 80 (2009),
Met., 51 (1999), No.5, 527-537. No. 4, 298-303.

4th Steeluniversity Annual Challenge

The winners of the World Steel Association 4th


steeluniversity.org Challenge were presented with the
trophy and cash prizes in Shanghai on March 19.
The 4th steeluniversity.org Challenge was held online in
November 2008 over a single 24 hour period. The entrants
were challenged to process an order of hot-rolled plate to
be used in an offshore wind turbine farm. The team
successfully producing the order at the lowest cost/tonne
within the allowed 24 hours was declared the winner.
The winning team of Yin Lu, Ge Hua and Chen Cen
from China’s Baosteel completed the production
Challenge at the lowest cost of $ 589,702. The runners-up
were Wang Chao and Li Shuguang from Ansteel who
came in with a cost of $ 589,810.
The winners of the Steeluniversity Challenge 2008 Award (from
The global competition attracted more than 16,000 left to right) Ge Hua, Yin Lu, Xu Lejiang (President, Baosteel),
attempts from 478 teams representing 26 different Chen Cen (all Baosteel), Ian Christmas (Director General, World
countries across all five continents. 95 teams succeeded in Steel Association), Wang Chao (Ansteel), Xu Hailiang
passing the whole challenge, which used a new simulation (representing Li Shuguang, Ansteel).
specially developed for this event.
China boasted the highest number of attempts, with increased significantly over the 4 years that the Challenge
large entries also seen from South Korea, Russia and India. has been held with teams praising the realism of the
The number of entrants and recorded attempts has modules and the experience they gain.

Integrated High Speed Video Installation at Corus

Imatek has recently installed its Integrated High Speed


Video system at the Corus Swinden Technology Centre
(STC) in Rotherham.
The new system will be used to improve understanding
of the performance of steel at high rates of strain and for
other general applications.
The Swinden Technology Centre of Corus in Rotherham
is the main R&D site in the UK focusing on product
research and applications research for the transport and
building & construction sectors.
Imatek’s Integrated High Speed Video system uses its
C3008 data acquisition system to combine signals
recorded during a test event together with high speed
video imagery, all under the control of Imatek’s ImpAcqt
software. Since the camera and the data acquisition share
the same trigger, data points and images can be precisely any of the Imatek range of impact testers or in a stand-
correlated. alone mode, making it a general purpose tool that can be
A video sequence provides a great deal of qualitative utilised by other parts of a customers operation and on a
information about the test event, and the software also wide range of test apparatus including UTM’s, pendulum
allows quantitative information to be extracted. and drop weight testers.
Integrated High speed video offers an advantage in As standard the High Speed Video option provides a
many application areas and is available as an option for complete package including C3008 data acquisition,

310 steel research int. 80 (2009) No. 4

You might also like