You are on page 1of 10

AISTech 2019 — Proceedings of the Iron & Steel Technology Conference

6–9 May 2019, Pittsburgh, Pa., USA


DOI 10.1000.377.195

Sub-Arctic Toughness From API Coiled Plate on a Steckel Mill

J. Hinton1, A. Harvey1, J. Lee1, M. Arafin2, L. Collins2 and K. Dunnett2


1
Primetals Technologies UK Ltd.
Sheffield Business Park, Europa Link, Sheffield, UK, S9 1XU
Phone: +441709723646
Email: john.hinton@primetals.com
2
Evraz Regina Steel.
100 Armour Road, Regina, Saskatchewan, Canada, S4P 3C7
Phone: +13069247695
Email: kendal.dunnett@evrazna.com

Keywords: API Grades, Through Process Metallurgy, Low Temperature Toughness

INTRODUCTION
High strength and toughness American Petroleum Institute (API) grade steels have to fulfil increasingly challenging
structure property demands in modern pipeline applications. New projects might require an increase in plate thickness to
allow for an upgrade in the operating pressure or superior toughness to combat reduced design temperatures i.e. for
installations in sub- arctic regions. Besides general measures of quality assurance such as dimensional tolerances, the
production of these grades is focussed on achieving specific contract properties. The principal mechanical tests are
tensile, Charpy Impact or Charpy V- notch (CVN) and the DWT (Drop Weight Tear) test. Table I summarises the
acceptance criteria for Product Specification Level 2 X70 and X80 welded pipe from API Specification 5L. Canadian
line pipe standard CSA Z245.1 would also apply depending on the contract.

Table I Overview of the key performance indicators from API Specification 5L [1]
YS (Rt0.5) TS (Rm) Ratio CVNa DWTTb
Grade MPa (psi) MPa (psi) Rt0.5/Rm
J (ft. lbf) Min. area %
Min. Max. Min. Max. Max.
485 635 570 760 27 to 68
X70 0.93 ≥ 85
(70300) (92100) (82700) (110200) (20 to 50)
555 705 625 825 40 to 81
X80 0.93 ≥ 85
(80500) (102300) (90600) (119700) (30 to 60)
a
CVN conducted on full specimen at 0°C (32°F); range of energy values is dependent on the specified outside diameter.
b
The average shear fracture area in the DWTT shall be ≥ 85% based upon a test temperature of 0°C (32°F).
The measurements of CVN and DWT tests are to ensure that brittle fracture does not occur during service. The CVN
impact test measures the energy absorbed during fracture, whilst the DWT test uses a full thickness specimen and
assesses the toughness via the appearance of the fracture surface. This provides an additional safety factor to guarantee
operating above the ductile to brittle transition temperature. The DWT test is generally accepted to have the best
correlation with full scale behaviour; with 85% shear area an accepted minimum requirement to ensure that the fracture
propagation mode is ductile.

© 2019 by the Association for Iron & Steel Technology. 1907


Depending on the end use application, it is common practice to specify the toughness requirements at an agreed lower
temperature than those shown in Table I. This requires the steel producer to make the final structure property
relationships the focus of the process route. One option is to identify a target microstructure to achieve the specification,
and then consider the upstream processing options to design the appropriate production stages to deliver this end goal.
This design will include the capacity of the installed equipment base (i.e. rolling mill power and torque), the
development of an appropriate time temperature strain path and the composition (within the API 5L limits) of the steel
grade. These requirements inform the steel making and rolling strategy. In practice, a tolerance range or process window
should be considered, and must be achieved with statistical certainty so that the process is repeatable and a production
route established. Therefore, to achieve the desired properties, excellent process control must be utilised in conjunction
with an appreciation for the underlying metallurgical relationships.
This paper reports production trials for low temperature API coiled plate undertaken on a Steckel Mill in North America.
The temperature goal for pipe properties was -45°C (-49°F). The target microstructure to achieve these in service
properties is described, followed by necessary process route considerations including the spiral pipe forming process.
Key process parameters, such as control of the Reheat Furnace (RHF) and the roughing temperature range are
discussed in more detail. The final section presents results from a series of production trials with significant ongoing
developments through the trials highlighted.

TARGET MICROSTRUCTURE
The phase transformation of austenite to ferrite during post rolling cooling can be used to develop different cooling
strategies to obtain a wide range of structure property relationships. A fine, uniform microstructure is the foundation for
a good combination of strength and toughness i.e. small grains or ‘packets’ of microstructure, small coherent precipitates
etc. [2, 3]. On cooling from austenite, the final microstructure can consist of a range of ferrite morphologies; ferrite,
pearlite (laminar ferrite + cementite), bainite and martensite. Each morphology can be associated with macro properties
that relate to the microstructure features; ranging from soft, but ductile ferrite to hard but brittle martensite.
For sub-arctic requirements, highly organised morphologies e.g. packets of ferrite plates that have the same
crystallographic orientation, may be deleterious to the mechanical properties. This is because in-service fracture or failure
processes can move readily through the microstructure, with individual plates within these packets, offering limited
resistance. One ferrite morphology rarely described as organised is acicular ferrite. This chaotic microstructure has a
‘basket weave’ appearance. ‘Acicular’ means shaped and pointed like a needle, which is the appearance when viewed
in two dimensions. The three dimensional grains have a thin, lenticular shape. The needle-like plates nucleate
heterogeneously on intragranular point nucleation sites, such as inclusions, and radiate in different directions. Cracks
propagating through the microstructure are believed to be frequently deflected. At a macro level, superior properties,
especially toughness are often realised.
The acicular ferrite transformation range is 400 to 600°C (750 to 1110°F) and in many respects is a similar
transformation to upper bainite. The microstructure morphologies differ because bainite sheaves grow from austenite
grain boundaries; whereas acicular ferrite platelets nucleate at intragranular nucleation (or point) sites, so that parallel
ferrite plates cannot readily form. Acicular ferrite is sometimes referred to as intragranular upper bainite [4].
To promote an acicular ferrite microstructure, the density of intragranular nucleation sites should be increased
compared to grain boundaries sites. This is not straightforward as inclusions or inhomogeneities are less effective in
nucleation compared to grain boundaries. A second option is to reduce the effectiveness of the austenite grain boundary
as a nucleation site either by alloying or by changing the finish rolling / cooling strategy to allow the nucleation of a
small amount of grain boundary ferrite. In both cases the aim is to disrupt significant grain boundary ferrite formation to
take advantage of the increased driving force for intragranular nucleation at lower temperatures. Given these
microstructure challenges, the likelihood of an acicular ferrite structure in the final plate material during production is
increased if the parameters summarised in Table II are considered.

1908 © 2019 by the Association for Iron & Steel Technology.


Table II Summary of Parameters to that should be considered to promote an Acicular Ferrite final microstructure.

Parameter Process Consideration

Increase defects within the • This informs the rolling strategy that there should be significant
austenite grain structure. deformation below the non-recrystallisation temperature (TNR).
• This increases the dislocation density, shear bands and mismatch
within the crystal structure; all of which have the potential to act as
acicular ferrite nucleation sites.
• This also helps breakdown the prominent recrystallised austenite
texture, and increase the likelihood of different orientations in the final
microstructure.

Increase the number and size of • It has been shown that inclusions rich in titanium are effective in
inclusions. producing acicular ferrite [4]. A specific titanium compound has not
been reported as the microanalysis is difficult and the data accuracy a
consideration.
• The heterogeneous nature of non-metallic inclusions is one of the
reasons why there is a lack of clarity in the nucleation mechanism for
acicular ferrite.

Tailoring the chemistry; must • Consider an increase in Nb; this solute is well known to have an effect
remain within the limits set by on the recrystallisation of austenite, raising TNR and providing
the API 5L Specification. inclusions (precipitates).

SPIRAL PIPE FORMING PROCESS


Before discussing the influence of the rolling mill configuration and processing parameters, a comment on the
influence of the final as-rolled microstructure during pipe formation is appropriate. To produce pipe, coiled plate
feedstock uses spiral pipe production methods. For the API 5L Specification, the mechanical properties and product
specification levels are considered in end use rather than original coiled Steckel plate. Mechanical testing will occur
after the spiral pipe forming process and specimen flattening. To this end, further complications arise for the steel
producer in being able to precisely define the as-rolled properties i.e. the appropriate structure property relationships in
order to achieve the final pipe requirements.
During spiral pipe forming and specimen flattening, the inner and outer walls of the pipe receive different strains; i.e.
tension – compression on the outer wall and compression – tension on the inner wall. Due to this varying strain history,
test specimens generally show lower yield strength than in the as-rolled coiled plate condition. This is variation is
understood as the Bauschinger effect [5] which refers to the phenomenon where the effect of strain hardening is reduced
when a material, already deformed in one direction, is deformed again in the opposite direction. The Bauschinger effect
is influenced by the underlying microstructure and the generation, migration and pile up of dislocations during tension
and compression. For spiral formed pipe, the Bauschinger and strain hardening effects are more dominant towards the
inner and outer surfaces and increase from the centre to the outer or inner side as the pipe forming strains increase.
Acicular ferrite/bainite microstructures are more likely to exhibit continuous yielding (high dislocation density with a
high mobility) whilst a ferrite/pearlite microstructure is more likely to exhibit discontinuous yielding [6, 7].
The yield strength is only one key performance indicator, the other being toughness. Given the increase in complexity
of predicting toughness, it is understandable that the effect of pipe forming is not well understood. Given that the yield
strength changes due to the Bauschinger effect, it is considered best practice to assume that the toughness decreases.
There is limited statistical data but as a general industrial rule it is expected that steel producers should target the
required toughness at temperatures 10 to 20°C below the in service pipe requirements.

ROLLING MILL CONFIGURATION


For API pipeline steels, cast slabs are reheated and processed into coiled plate or discrete plates using various mill
configurations. Plate Mills, Steckel Mills, Hot Strip Mills, Endless Strip lines represent a range of mill configurations.
In general, the process route flexibility afforded to the steel producer decreases in the order that the configurations are
stated. Plate Mills produce shorter, wider, discrete plates with rolled lengths typically limited to less than 40 to 50m. The
other Mill configurations can produce coiled plate which is typical feedstock for the spiral pipe forming described earlier.
In modern or revamped mill lines, all of these mill arrangements are followed by various types of accelerated cooling

© 2019 by the Association for Iron & Steel Technology. 1909


systems with the most intense configurations being able to achieve cooling rates of ~ 30 to 35°C/s for 25.4mm (1”)
product.

Plate Mill Steckel Mill

Figure 1Schematic diagram of a finishing stand on Plate Mill and a Steckel Mill.
As shown in Figure 1, Plate Mills are typically single or two stand reversing mills whilst Steckel Mills are reversing
mills with the addition of heated coil furnaces on the entry and exit sides of the mill. This enables the effects of
temperature run down to be lessened and the production of products with similar lengths to those on a Hot Strip Mill is
possible. Coils up to 25.4mm (1”) thick are now routinely produced. The mill layout itself acts as one constraint on the
development of time temperature strain paths for new products. One of the benefits of a Steckel Mill compared to a
Plate Mill is the single piece weight that can be rolled by producing coiled product. However unlike a Hot Strip Mill,
where the finishing sequence occurs in a single rolling direction, a Steckel Mill involves reversing and coiling as the
product thickness is reduced. This leads to an increase in processing time. Compared to a Plate Mill, typical finishing
times maybe greater than twice as long in a Steckel Mill for products of comparable thickness and width. In addition, the
increased product length leads to inhomogeneity in the time temperature strain paths from the head to tail. The
implication of this characteristic of Steckel Mill rolling on the final product performance must also be considered by the
steel producer.

ALLOY DESIGN AND ROLLING PARAMETERS


As a platform to achieve the target acicular ferrite microstructure, the alloy design and key process parameters should
be established to attain a fine uniform cross sectional grain size in the as-rolled coiled plate. The refinement of the as-
cast austenite grain structure provides the basis for the final metallurgy; i.e. the desired yield/tensile strength, elongation
and high toughness at low impact temperatures. A detailed summary of the key stages with recommendations to
consider is provided in Table III. From this table, the influence of two factors will be considered in more detail to
achieve superior low temperature toughness. One is control of the RHF and the second is the temperature range for the
roughing sequence, in particular the last pass roughing temperature.

Control of Reheat Furnace (RHF) temperature


In industrial RHF operations, the uniformity of slab reheating along the length and through thickness is a primary target.
The second is to maximise the solubility of the micro-alloyed elements so that they can be effective downstream i.e. for
recrystallisation kinetics, grain refinement and precipitation. This requires good control of the target reheat temperature;
one that allows significant dissolution of existing precipitates whilst keeping the temperature optimum to avoid
abnormal austenite grain growth. The through thickness microstructure and compositional inhomogeneity will
influence the downstream processing and potentially reduce the impact toughness and DWTT performance.
The low solubility of titanium nitrides has led to them being used to control austenite grain growth. As the RHF
temperature increases, the double effect of increasing the likelihood of precipitate dissolution and precipitate
coarsening should be considered [8] i.e. there are less TiN particles and those that are present are larger in size. For
superior low temperature toughness this is considered detrimental; a potentially inhomogeneous microstructure and
coarser TiN particles that can act as stress concentration points. The focus should therefore be on a less than
stoichiometric Ti/N ratio [8, 9] which will maintain fine, coherent TiN precipitates. Figure 2 compares the final grain
size distribution from two API coiled plates form the production trials reported later. Due to concerns about the through
thickness compositional inhomogeneity and the full dissolution of Nb(CN)s, the RHF temperature in the first production

1910 © 2019 by the Association for Iron & Steel Technology.


trial was approximately 50°C higher than in the second production trial. With comparable TiN ratios and downstream
processing, the grain size distribution was found to be larger for the higher RHF temperature (Figure 2a vs 2b). The
inhomogeneity in the final microstructure is characterised by the significant population of grains (> 10%) greater than
20μm.

Figure 2a Figure 2b

0 10 20 30 40 0 10 20 30 40

Figure 2 A comparison of the grain size distribution from two API coiled plates; Figure 2a with higher RHF temperature
and Figure 2b with a lower RHF temperature.

Table III Summary of the Key Through Process Parameters for coiled API Steckel plate.

Process Overview
Parameter

1) Alloy Design • A C-Mn-Si based composition is a nominal starting point for alloy design.
• Solid solution strengthening (Mn, Ni and Mo) and micro-alloying (Nb, V and Ti).
• Nb is an important element in designing the rolling strategy i.e. determining TNR
which in turn determines the temperature ranges for roughing, finishing and the
hold period.
• Nb(CN) to precipitate during the hold, finish rolling and cooling to the aid the
nucleation of a fine final microstructure.
• Due to the longer finish rolling times, coiled API Steckel plate is expected to
require a richer composition with increased micro-alloying. The longer time at
temperature coarsens precipitates which reduces their effectiveness. 30 to 50% more
Nb is expected than in plate chemistries.
• Ti (an element with a higher affinity for N than Nb) is added to fix N and potentially
allow a greater Nb % to dissolve during reheating. Potential to aid formation of
acicular ferrite.

2) As-cast Slab/ • For API high strength with good low toughness, a total reduction ratio (slab to
Structure plate thickness) of 10:1 is recognised as a minimum requirement.
• The other main targets are tight and consistent chemistry control, good internal
cleanliness and minimisation of downstream influence of the centreline.
• Good slab surface quality and dimensional control promote more efficient
downstream processing.

3) Reheat • Uniform heating.


Furnace • Aim to dissolve Nb into solution to precipitate again during the rolling process.
• TiN and AlN precipitates to control austenite grain growth.

© 2019 by the Association for Iron & Steel Technology. 1911


4) Roughing • Good consistent deformations through the sequence, increasing to a last pass
Sequence maximum reduction prior to the Hold Period.
• Break down of large columnar cast grains via recrystallisation in successive
passes; to ensure a fine-grained recrystallised microstructure.
• Typically expect > 60% reduction in roughing.

5) Hold Period • Pause in the rolling to allow cooling to below TNR and TST (recrystallisation stop
temperature).
• Nb(CN) precipitates form, pinning the austenite grain boundaries to prevent
recrystallisation on the resumption of rolling.

6) Finishing • The restart temperature will depend on mill capability (force, torque) for the given
Sequence product dimensions. Aim for the lowest temperature possible.
• Maximum deformations will also depend on mill capability (force, torque) for the
given product dimensions and temperature range.
• This should be designed to achieve good reductions per pass; > 15%.
• Deform and pancake the pinned austenite grains; creating dislocations and shear
bands in conjunction with precipitates, within the austenite grains to provide
nucleation sites for acicular ferrite.
• Typically expect > 70% reduction in Finishing.

7) Cooling • To target preferred acicular ferrite. A fine or very fine form of the ferrite with
Strategy small α- ferrite laths between carbon rich partitions.
• Target cooling rate will depend on composition and thickness.
• Expected to be between 20 and 30 °C/s.

Last Pass Roughing Temperature


Given the hereditary nature of microstructure, there is a strong relationship between the austenite and ferrite grain sizes;
i.e. a fine austenite grain structure can transform into fine ferrite morphologies on cooling. The roughing sequence
focuses on the breakdown of the as cast columnar grain structure and should target good, consistent reductions which
increase towards the hold period. Austenite grain refinement is achieved via successive static recrystallisation events
between reductions, with increasing strains per pass designed to achieve through thickness penetration. In the design of
this rolling strategy, the last roughing pass represents the last opportunity to refine the austenite grain size by static
recrystallisation. Due to long range grain boundary motion, recrystallisation is a thermally activated restoration
mechanism that returns the microstructure to an undeformed state. The recrystallised grain size is a function of strain,
strain rate and deformation temperature i.e. the recrystallised grain size reduces with increasing strain and reduced
deformation temperature. As the temperature decreases, the time to full recrystallisation (taken as 95% recrystallised
fraction) increases to the point where, due to thermal energy, the mechanism stops within a meaningful time frame. At
the same time the propensity for the precipitation of Nb(CN) increases as the temperature decreases. It is well-known that
the onset of precipitation delays recrystallisation. This is a function of the temperature, the concentration of Nb in
solution after the correct RHF practice and the strain per pass. Therefore, recrystallisation versus strain induced Nb
precipitation is a key phenomenon that informs the end temperature of the roughing sequence.
For TNR, there are a number of empirical equations published in the literature which take into consideration the micro-
alloying content as well as C and N [10, 11]. A more complete description is provided where the kinetics of strain
induced Nb precipitation are also taken into account. In this calculation, the recrystallisation kinetics (start and stop;
5% and 95% recrystallisation) and the time required for 5% precipitation relative to the equilibrium are determined [12,
13, 14]. An example calculation is shown in Figure 3. The dashed grey lines show the modelled start and finish
recrystallisation times, whilst the black dashed lines report the predicted start of Nb(CN) precipitation (5%). The
straight lines show empirical relationships for TNR. Point (1) defines TNR; 95% recrystallisation intersecting with the
modelled 5% Nb precipitation curve. The presence of Nb precipitates retards recrystallisation before it is complete.
Below this temperature there is the potential to develop a mixed microstructure (i.e. partial recrystallisation). Point (2),
the intersection of the 5% precipitation curves with 5% recrystallisation defines TST, the stop recrystallisation
temperature. Below this temperature recrystallisation does not start. This is the temperature at which the deformation
conditions and composition combine to completely stop recrystallisation; i.e. the potential pancaking of austenite grains
has begun. Figure 3 shows that for this combination of process parameters (composition, deformation) there is in
general a good agreement between the empirical TNR and TST equations and the predicted start of Nb(CN) precipitation.

1912 © 2019 by the Association for Iron & Steel Technology.


(1)

(2)

Figure 3 Prediction of non – recrystallisation (TNR) and stop – recrystallisation (TST) temperatures to target the end
roughing temperature; strain 0.25, strain rate 3s-1.
For a good uniform final microstructure from all micro-alloyed rolling strategies, it is generally recommended that the
final deformation temperature in the roughing sequence is greater than TNR whilst the entry temperature for the finishing
sequence (i.e. after the hold period) is below TST. This avoids the mixed microstructure region if deformations occur
between TNR and TST, which is potentially detrimental to final mechanical properties. For superior low temperature
toughness, the last roughing pass should be as close to TNR as is practically possible. This allows full recrystallisation,
but due to subsequent Nb precipitation, the extent of austenite grain growth in the hold period is reduced. From Figure
3 a target last pass roughing temperature would be in the range 1020 to 1050°C (1870 to 1920°F).
A further practical consideration is that Figure 3 reports isothermal conditions. After drop out from the RHF, the
workpiece develops a thermal gradient as the surface is cooled during contact with the rolls and surrounding
environment. Due to the large hold ratio (> 70% reduction expected in finishing) the thermal gradient from surface to
centre is likely to be greater than 50°C (90°F). With a good, fine, uniform structure integral to low temperature
toughness, the steel producer should consider the thermal gradient when designing the temperature range of the
roughing sequence.

RESULTS FROM PRODUCTION TRIALS


This section reports the progress of industrial production trials that were supported by Primetals Technologies UK Ltd.
and designed to deliver superior low temperature X70 spiral welded pipe from coiled 19.1mm (¾”) API Steckel plate.
The aim was to achieve the final microstructure and properties via a process route and parameters that considered the
statements in Table III. The challenges of translating the desired process variables into full scale successful production
trials, with all the through process and multi-variable complexity were evident. Steel producers have their own standard
melt practices, rolling practices etc. that control and ensure repeatability of their core products. If the production trial
represents a departure from this, then it is reasonable to expect that some degree of learning will be required during the
trials. The focus of the results presented here is the DWTT performance.
The DWTT results (including standard deviations) for the first two production trials are shown in Figure 4a which
reports percentage shear fraction with decreasing test temperature. For both trials a close to conventional X70
composition was used. The results from Trial 1 are show a typical transition from ductile to brittle fracture as the DWTT
temperature decreases. The test results were good to approximately -20°C (-4°F) with greater than 85% shear fraction
reported. Below this test temperature, the transition to a more brittle fracture became apparent as did the variability in
the results of the DWT tests. Following the results from Trial 1, the process parameters and temperatures were
reviewed. Trial 2 was designed based on maintaining the composition but with a focus on lowering the process

© 2019 by the Association for Iron & Steel Technology. 1913


temperatures at a number of key stages in the process route. This included the RHF temperature to reduce the propensity
of abnormal austenite grain growth and the final roughing temperature to achieve a heavy last pass deformation as close
to TNR as possible (as discussed in the previous section). At the end of rolling, the coiling temperature was also
decreased to ensure the cooling interrupt temperature was lowered. The results for Trial 2 (Figure 4a) report improved
DWTT performance at all test temperatures. The improvements were particularly noticeable at -30°C (-22°F). This was
due to the greater grain refinement and a finer final microstructure as a result of the colder processing temperatures. The
performance was considered borderline for the minimum 85% shear fraction requirement and the variability of the DWT
test became more noticeable as the test temperature decreased.

Figure 4a Figure 4b

Figure 4 Comparison of DWTT results from a) Trials 1 and 2 and b) Trials 2 and 3. The effects of processing temperature
and composition can be identified.
Detailed microscopy from the first two production trials found that centreline segregation was potentially an issue; the
hardness of the centreline was approximately 20 points harder than at a ¼ thickness. A new trial was planned with a
modified X70 composition, one that was designed to reduce the effect of the centreline segregation. The ‘Low’
temperature parameters known to achieve the improvement in Figure 4a were retained. Figure 4b reports percentage
shear fraction with decreasing DWTT temperature for the new trial. The results show improvements in the percentage
shear fraction at all test temperatures and increase the average percentage shear fraction at -45°C (-49°F) to
approximately 70%. From these trials there was improved product consistency, with greater than 90% shear fraction
reported with more consistency at temperatures down to -30°C (-22°F).
In reviewing the trial process data the roughing sequence was identified as an area for possible improvement. This was
an increase to the final pass reduction to 25%. Due to potentially coming close to the limits of the Roughing Mill, a
product with a narrower width was trialled. The results are shown in Figure 5, which also includes the DWTT curves
for the earlier production trials for comparison. Similarly to the previous trials with the modified X70 composition, an
improvement was reported at all DWTT temperatures. The greatest improvement is shown at -45°C (-49°F), with the
shear fracture percentage around 90%. This indicates that the energy shelf has been moved to lower test temperatures.
Future trials are being planned to increase the final product dimensions and further improve mechanical property
performance.

1914 © 2019 by the Association for Iron & Steel Technology.


Figure 5 Comparison of Comparison of DWTT from production trials 2, 3 and 4. The improvement in percentage shear
fraction with DWTT temperature for each trial is shown.

CONCLUSIONS
Targeting sub-arctic toughness performance from coiled 19.1mm (¾”) API X70 Steckel Plate represents a challenge for
the steel producer. It represents an example of a premium product where the through process metallurgical
understanding is paramount, both in selecting the target structure property relationships but then in delivering this
outcome from a complex multi-stage processing route. The final pipe performance can be traced upstream through all
stages to the initial steel making and casting. Once a slab is cast, the remaining process parameters need to be tightly
controlled and repeatable throughout a production run. The production trials were designed and performed by Evraz
Regina Steel with support from Primetals Technologies UK Ltd. The results in the final section show a clear
chronological improvement in product performance with consistent DWTT performance at a test temperature of -30°C
and clear progress towards the target of -45°C (-49°F). This represents a development in the capability of the installed
equipment base, the process consistency and in the understanding of the key through process metallurgical parameters.

REFERENCES
1. API Specification 5L, 45th Edition, API Publishing Services, Washington DC, 2012, pp32.
2. E.O. Hall, “The Deformation and Ageing of Mild Steel”, Proc. Phys. Soc., Sec. B, Vol. 64, 1951, pp. 747 – 753.
3. N.J. Petch, “The cleavage strength of polycrystals”, Journal of Iron and Steel Institute, Vol. 174, 1953, pp. 25 – 28.
4. R.W.K Honeycombe and H.K.D.H Bhadeshia, “Acicular Ferrite”, Steels; Microstructure and Properties, 3rd edition,
Elsevier, Oxford, UK, 2006, pp155.
5. G.E. Dieter, Mechanical Metallurgy, 3rd edition, McGraw – Hill Book Co., New York 1986, pp236.
6. C.W. Choi, H.J. Koh and S. Lee, “Analysis and Prevention of Yield Strength Drop during Spiral Piping of Two
High-Strength API-X70 Steels”, Metallurgical and Materials Transactions A, Vol. 31A, 2000, pp2669.
7. S.S. Sohn, S.Y. Han, J. Bae, H.S. Kim and S. Lee, “Effects of microstructure and pipe forming strain on yield
strength before and after spiral pipe forming of APIX70 and X80 linepipe steel sheets”, Materials Science &
Engineering A, Vol. 573A, 2013, pp. 18 – 25.
8. S.F. Medina, M. Chapa, P. Valles, A. Quipse and M.I. Vega, “Influence of Ti and N contents on Austenite Grain
Control and Precipitate Size in Structural Steels”, ISIJ International, 1999, Vol. 39, No. 9, pp. 930 – 936.

© 2019 by the Association for Iron & Steel Technology. 1915


9. M. Ohno, C. Murakami, K. Matsuura and K. Isobe, “Effects of Ti Addition on Austenite Grain Growth during
Reheating of As-Cast 0.2 mass % Carbon steel”, ISIJ International, 2012, Vol. 52, No.10, pp.1832 – 1840.
10. F. Boratto, R. Barbosa, S. Yue and J.Jonas, “Effect of Chemical Composition on Critical Temperatures of
Microalloyed Steels”. International Conference on Physical Metallurgy of Thermomechanical Processing of Steels
and Other Metals, THERMEC ‘88, 1988, Vol. 1, pp383 – 390.
11. 11 D. Bai, R. Bodnar, J. Ward, J. Dorricott and S. Saunders, “Development of Discrete X80 Line Pipe Plate at
SSAB Americas”, Proceedings of AISTech 2011, International Symposium on the Recent Developments in Plate
Steels, Warrendale, 2011, pp. 13-22.
12. R. Abad, A.I. Fernández, B. López and J.M. Rodriguez-Ibabe, “Interaction between Recrystallization and
Precipitation during Multipass Rolling in a Low Carbon Niobium Microalloyed Steel”, ISIJ International, 2001,
Vol. 41, No.11, pp. 1373 – 1382.
13. P.D. Hodgson and R.K. Gibbs, “A Mathematical Model to Predict the Mechanical Properties of Hot Rolled C-Mn
and Microalloyed Steels”, ISIJ International, 1992, Vol. 32, No. 12, pp. 1329 – 1338.
14. B. Dutta, C.M. Sellars, “Effect of composition and process variables on Nb(C, N) precipitation in niobium
microalloyed austenite”, Materials Science and Technology, 1987, 3, pp. 197 – 206.

1916 © 2019 by the Association for Iron & Steel Technology.

You might also like