J. FODER et al.: MEAN-FLOW-STRESS ANALYSIS OF LABORATORY HOT-ROLLED S1100QL STEEL ... 901–908 MEAN-FLOW-STRESS ANALYSIS OF LABORATORY HOT-ROLLED S1100QL STEEL WITH MINOR Nb ADDITION MFS ANALIZA LABORATORIJSKO VRO^E VALJANEGA JEKLA S1100QL Z MANJ[IM DELE@EM Nb Jan Foder 1 , Grega Klan~nik 1* , Jaka Burja 2 , Samo Kokalj 1 , Bo{tjan Brada{kja 1 1 RCJ d.o.o., Cesta Franceta Pre{erna 61, 4270 Jesenice, Slovenia 2 Institute of Metals and Technology, Lepi pot 11, 1000 Ljubljana, Slovenia Prejem rokopisa – received: 2020-07-06; sprejem za objavo – accepted for publication: 2020-10-05 doi:10.17222/mit.2020.125 Laboratory hot-rolling of ultra-high-strength martensitic steel S1100QL was carried out by measuring the material resistance during hot-rolling using rolling force measurements. A simplified approach was used to calculate the mean flow stress as an in- direct method to analyse the material’s response during hot rolling. It was shown that the method gives satisfactory results, re- garding the rough detection of different temperature-deformation regions like crossing the region of considerable solute drag and precipitation hardening, known as the non-recrystallization region. The transition from the recrystallization region, where softening mainly by static or meta dynamic recrystallization is replaced with a non-recrystallization region having austenite pancaking, is clearly distinguishable for given rolling parameters and final strip thickness. The grain size distribution of the S1100QL was also evaluated from the as-cast state to the heat-treated condition. Keywords: S1100QL, micro-alloying, hot-rolling, CALPHAD Izvedeno je bilo laboratorijsko vro~e valjanje ultra-visoko-trdnostnega martenzitnega jekla S1100QL z analizo materialnega odziva na preoblikovanje preko meritev sil valjanja. Izra~unana je bila srednja napetost te~enja (MFS), preko katere lahko indirektno analiziramo odziv materiala med vro~im valjanjem. Dokazali smo, da z uporabo MFS analize dobimo zadovoljive rezultate, vezano na detekcijo razli~nih temperaturno-deformacijskih podro~ij, kot je obmo~je kjer raztopljeni elementi povzro~ajo dodaten upor in obmo~je izlo~evalnega utrjevanja, znano kot obmo~je ne-rekristalizacije. Za dane parametre valjanja in kon~no debelino traku je jasno viden prehod iz podro~ja, kjer poteka meh~anje s stati~no ali metadinami~no rekristalizacijo, v podro~je kjer je ta omejena oziroma ne poteka, s prehodom avstenitnega kristalnega zrna iz ekviaksialne v pala~inkasto obliko. Dolo~ena je bila tudi razporeditev velikosti prvotnih avstenitnih kristalnih zrn od litega do kon~nega toplotno obdelanega stanja jekla S1100QL. Klju~ne besede: S1100QL, mikrolegiranje, vro~e valjanje, CALPHAD 1 INTRODUCTION Ultra-high-strength steels (UHSS) and high-strength low-alloyed steels (HSLA) have a high strength-to- weight ratio and are often used in the automotive indus- try, construction, mining and elsewhere. HSLA and UHSS are both used for demanding welded structures, with the emphasis on achieving a high yield strength and ductility. Various strengthening mechanisms like grain refinement, the formation of deformation substructures (dislocations), solid-solution strengthening and precipita- tion strengthening are used to achieve the ultra-high strength. The strengthening mechanism depends on the given alloy composition and the production route. In this paper the influence of hot rolling is emphasized. Never- theless, a combination of secondary metallurgy and cast- ing practice with final heat treatment is also of great im- portance for consistent production quality. Therefore, similar chemical compositions and different technologi- cal routes, such as different rolling schedules, will un- doubtedly lead to different material properties, which will also be measured in the final heat-treated condition. Unfortunately, all strengthening mechanisms, with the exception of grain refinement, have a negative effect on the impact toughness and ductility. Hot rolling is there- fore a key step in the production of UHSS and HSLA with superior properties. 1.1 S1100QL Niobium microalloyed steel S1100QL (Mat.No 1.8942) is a non-standardized fine-grained, fully marten- sitic UHSS. S1100QL must meet the minimum yield strength of 1100 MPa (with a tensile strength between 1200 MPa and 1500 MPa) and a Charpy impact tough- ness of at least 27 J transverse and 30 J longitudinal to the rolling direction at –40 °C. 1 Grain size control is achieved with proper hot-rolling and selected reheating temperatures. Low-temperature tempering is performed after water quenching to maintain adequate yield strength and increase in ductility. This paper focuses on achieving the minimum yield strength of 1100 MPa by quenching and low-temperature tempering. In the case of on-line heat treatment, austenite conditioning prior to quenching must be considered. Irrespective of the type of Materiali in tehnologije / Materials and technology 54 (2020) 6, 901–908 901 UDK 621.77:620.1:669.1 ISSN 1580-2949 Original scientific article/Izvirni znanstveni ~lanek MTAEC9, 54(6)901(2020) *Corresponding author's e-mail: klancnik.grega@gmail.com (Grega Klan~nik) heat treatment, it is recognized that the martensite block and packet size tend to scale linearly with the parent aus- tenite grain size (PAGS), affecting the size of the high-angle substructures (>15° misorientation). 2 This is important if the packet size is considered as the "effec- tive grain size" in relation to the yield strength and toughness of the lath martensite. 3 The effective grain size and distribution in combination with steel cleanliness, which is the result of secondary metallurgy practice, de- termines the functional properties such as the weldability and formability of high-strength steels, therefore an opti- mal rolling schedule is important and necessary. 4,5 Nowa- days, the advanced hot-rolling technology for HSLA and UHSS plate production is based not only on austenite grain refinement and control over the precipitation of ni- trides, carbides and carbo-nitrides, but also on transfor- mation control during cooling after hot rolling with the possibility of significantly reducing the alloying, such as reducing the expensive molybdenum, and still obtaining excellent or even improved mechanical properties. This is possible if the process of controlled hot rolling is cou- pled with on-line controlled cooling using accelerated cooling (ACC), ultra-fast cooling (UFC) or even direct quenching (DQ) technology. 6 This promising technology of thermo-mechanical control processes (TMCP) with controlled cooling, not only to improve the strength and ductility properties for certain steel grades, but also for steel mill time and energy efficiency, still needs to be adapted for HSLA and UHSS production in the local re- gion. 1.2 Strip/plate rolling Phenomena occurring during the hot deformation (hardening, softening etc.) of steels are usually studied by hot compression, tension and torsion tests on thermo-mechanical simulators. 7–9 Numerous specimens and tests are required to obtain the relevant data on the hot-deformation behaviour. The use of a laboratory roll- ing mill for the evaluation of the mean flow stress (MFS) is an alternative method for the investigation of the hot deformation of steels. It is particularly useful for obtain- ing important data during successive reductions in the hot-rolling process. In addition, basic mechanical proper- ties such as strength and toughness, bending, etc. are ob- tained from the sample in a similar way to industrial strip/plate production, as larger samples are produced and tested. The final grain size can be achieved through strain- free austenite, formed by plastic deformation. Only lim- ited grain refinement is expected after cooling and the transformation from austenite to ferrite. This means that during conventional hot rolling and recrystallization-con- trolled rolling (RCR), the main deformations are carried out as part of the roughing phase. The latter with an em- phasis on higher per-pass reductions. When finishing passes are introduced, they are mainly adjusted to the target dimension tolerance and flatness in relation to the selected finish rolling temperatures (FRT). A finer grain size is expected when additional deformation at lower temperatures is introduced to obtain deformed austenite grains with deformation bands and strain-induced precip- itates (SIP). Combined two-stage rolling of roughing and finishing with a certain time delay between the roughing and finishing phase is called controlled rolling. 10 Accord- ing to Dutta et al., 11 the recrystallization limit tempera- ture (RLT) is the lowest temperature for the roughing phase to prevent the formation of a mixed microstructure by partial recrystallization. A mixed microstructure causes the scattering of the final mechanical proper- ties. 5,10 The RLT temperature can be interpreted as the non-recrystallisation temperature T nr , due to limited recrystallisation. Based on the recrystallization behaviour, plastic de- formation can be divided into three main characteristic regions (Figure 3): 5,10 • Region I – typical for deformation above the RLT, the temperature where approx. 95 % recrystallization still occurs. The deformation per pass and the interpass times control the softening ratio achieved by static recrystallization (SRX) or meta dynamic recrystal- lization (MDRX) during rough rolling or RCR. • Region II – is below the recrystallization stop tem- perature (RST), the temperature below which less than 5 % recrystallization still occurs. • Region III – is between region I and II (between RLT and RST) and should be avoided due to the limited control of recrystallization, grain growth and precipi- tation. It is also referred to as the partial recrys- tallization region (95 % to5%ofSRX). Below the RST, where recrystallisation of less than 5 % is expected, both finish rolling and hot levelling are usually carried out for controlled rolling. In some cases, due to the sufficient number of passes and associated temperature drop of the final passes, this is also done by conventional hot rolling of thin plates or strips. Softening at low temperatures can be influenced by the introduc- tion of microalloying elements such as niobium for SIP. According to Dutta and Sellars, 11 the retarding effect of niobium on austenite recrystallization is observed to a lesser extent for niobium in a solid solution and more pronounced with the SIP of Nb(C,N) from the austenite matrix, which means that sufficient accumulation of the deformation for SIP can effectively stop recrystallization recognized as Region II (below the RST). It has also been recognized that RST occurs when the time for 5 % recrystallization (or less) is equal to the time for5%of precipitation of Nb(C,N). 11 Inhibition of recrystallization is effective until the end of precipitation and loss of the effective pinning force (Zenner pinning). Austenite grains are therefore elongated (pancaked) in rolling di- rection if sufficient strain is accumulated and that mini- mum incubation time is respected for the successful SIP of Nb(C,N) (see Figure 1c). 12 It is also recognized that deformation accelerates precipitation, providing prefer- J. FODER et al.: MEAN-FLOW-STRESS ANALYSIS OF LABORATORY HOT-ROLLED S1100QL STEEL ... 902 Materiali in tehnologije / Materials and technology 54 (2020) 6, 901–908 ential sites for the nucleation and precipitation of Nb(C,N) at lower supersaturation, as in the case of the undeformed austenite 11 and with that expected change in flow stress behaviour (sometimes written as k f ), as it is influenced by the introduction of the dislocation density variation, , inside the grains: 11,13 = Gb (1) Where is a constant, G is the shear modulus and b is the Burgers vector. The dislocation density is not so trivial to measure, 14 so instead of observing metallurgical phenomena associated with the change in dislocation density, a recognized change of grain-boundary migra- tion due to SIP is possible with various methods, but also by measuring the change in the average flow stress. 15 1.3 Alloy design Alloy design is crucial for the rolling schedule design and final microstructure that is achieved after heat treat- ment. The chemical composition of S1100QL with mi- nor niobium and titanium additions is summarised in Ta- ble 1. In the case of niobium and titanium carbo-nitride forming, they act as a retarding force on grain-boundary migration. This is also important for limiting the grain growth during cooling. In addition to Zenner pinning with rather moderate niobium and titanium contents, grain-boundary mobility can be influenced by the solute drag of niobium, titanium and also molybdenum. 2,16 Tita- nium and niobium additions are minor, which excludes the possibility of coarse nitride and carbonitride forma- tion already from the molten steel, see Figure 1 a and Figure 4. 17 Niobium has an important effect on RST and RLT, ei- ther in solution or even more so as strain-induced Nb(C,N) precipitates. 11,15–21 As mentioned above, nio- bium in solid solution segregates to austenite grain boundaries, which causes drag and thus retards the recrystallization by the solute-drag effect. 16,18 Nb(C,N) precipitates are more effective in retarding recrystal- lization. They nucleate first at the austenite grain bound- aries (RLT) and later at the dislocations with matrix pre- cipitation (RST), pinning them and thus retard recrystallization by preventing grain-boundary migra- tion. 19–21 The non-recrystallization temperature (T nr )i sn o r - mally used to determine the critical temperature at which strain accumulation occurs during hot rolling. T nr is nor- mally assumed to be similar or close to the RLT, al- though the exact position depends on other variables such as strain, strain rate, experimental method used, etc. This means that T nr can be identified between RLT and RST, depending on the experimental method and the given parameters. According to literature 5,10 , the MFS range for Region III in standard HSLA grades during hot rolling with typical per pass reductions and strain rates for thin plate or strip rolling is between 150 MPa and 200 MPa. To correlate the measured MFS change in cor- respondence to T nr , we have used the widely accepted Boratto’s equation for HSLA grades: 22 T nr = 887 + 464C + 6445Nb – 644 Nb + 732V – 230 V+ 890Ti + 363Al – 357Si (2) where all the elements are given in mass percent. 1.4 Rolling load model According to Sims 23 , the calculation of rolling force F can be easily calculated by: F = k fm bl’ d Q f (3) where b is the plate width, l’ d is the horizontal projec- tion length of the contact arc between the squashed roller and plate, Q f is the geometric factor and k fm is the mean metal deformation resistance or the mean value of constrained yield stress in the roll gap. It is also possible to introduce the effect of the rolling force by tensile stress (i.e., Steckel). 24 It is important that the equations can also be used for strain-hardening materials, not only for ideal plastic-rigid material. If plane deformation is assumed, some precautions must be taken if the width/thickness ratio per pass is less than 5. In this pa- per an additional simplification has been made by set- ting Q f to unity. 2 EXPERIMENTAL PART The experimental S1100QL steel was produced in a vacuum induction furnace and cast as a 12-kg ingot (80 × 80) mm 2 . The ingot was heated to 1150 °C and soaked for 2 h before hot forging into a (60 × 60) mm 2 billet. The billet was cut into two parts. Individual billets intended for strip hot rolling were then reheated to 1200 °C for 30 min (a total 2.5 h for temperature homo- geneity) to ensure that the niobium was completely in the austenite solution in order to achieve an effective and uniform recrystallization retardation by solute drag or SIP during hot rolling. The solubility temperature for Nb(C,N) was determined to be 1122 °C using the Irvine 25 equation. An additional infrared camera Optris PI 1ML with a working temperature range between 450 °C and 1800 °C was used to pre-define the tempera- ture distribution measurement before and during rolling at interpass times. The laboratory billet was then re- J. FODER et al.: MEAN-FLOW-STRESS ANALYSIS OF LABORATORY HOT-ROLLED S1100QL STEEL ... Materiali in tehnologije / Materials and technology 54 (2020) 6, 901–908 903 Table 1: Chemical composition of S1100QL billet, in mass fractions, (w/%) CS iM nC rN iM oA lT iN bBNF e 0.18 0.23 0.87 0.52 1.3 0.41 0.06 0.013 0.014 0.0011 0.0031 bal. heated as described above and hot rolled in a predefined 11 passes to a final thickness of 12 mm, using a single stand mill with subsequent continuous air-cooling. The soaking temperature and the number of passes with the specified per-pass reduction made it possible to achieve a low-temperature finish. Hot rolling was performed with- out interruption or time delay between the roughing and finishing to observe the MFS behaviour in the region of the RLT and/or RST (defined as T nr in this paper) for the given chemical composition and related rolling parame- ters. The rolling loads were continuously measured for indirect MFS evaluation by stress measurements on the rolling mill. The last step in the production process was a laboratory heat treatment by re-austenitization of the samples in a temperature-controlled furnace using addi- tional thermocouples (Type-K) for the samples’ austenit- ization control. Thus, after austenitization at 900 °C and a holding time of 20 min, the steel was first water quenched, followed by low-temperature tempering at 200 °C and a holding time of 60 min. The general chemical composition was determined by optical emission spectroscopy (OES), ARL MA-310. Carbon and sulphur were determined by the combustion method using a LECO CS-600 and nitrogen using a LECO TC-500, Table 1. The base composition also cor- responds to other grades with a lower yield strength ac- cording to standardized steel qualities defined in reference 26 . The microstructure and PAGS evolution of S1100QL in the as-cast, hot-forged, hot-rolled and heat-treated conditions was characterised, using scanning electron microscope with field emission gun (FE-SEM), Zeiss Supra VP55. Samples were prepared following a stan- dard metallographic procedure of grinding, polishing and etching using 2 vol. % Nital solution. PAGS and distribu- tion were determined using imageJ software by analys- ing an average of 400 grains. Thermodynamic prediction, using Thermo-Calc 2020a, was also conducted using TCFE7 and TCFE9 da- tabases for the estimation of the thermodynamically sta- ble characteristic temperatures for carbo-nitride precipi- tation. A simplified isopleth equilibrium Fe-C phase diagram (Figure 4) was calculated to provide a better vi- sual representation of the temperature dependence of the titanium- and niobium-rich precipitates with respect to the carbon variation. To evaluate the mechanical properties in the final tempered condition, longitudinal tensile specimens were prepared and tests were carried out according to EN ISO 6892-1:2017 B60 using Zwick/Roell Z600. 27 Standard V-notch specimens for the Charpy pendulum impact test were machined according to EN ISO 148-1:2016. 28 Im- pact toughness should be higher than 27 J for transverse test pieces, tested at –40 °C, as specified in the standard EN 10025-6:2019 for similar HSLA grades. 26 J. FODER et al.: MEAN-FLOW-STRESS ANALYSIS OF LABORATORY HOT-ROLLED S1100QL STEEL ... 904 Materiali in tehnologije / Materials and technology 54 (2020) 6, 901–908 Figure 1: Secondary electron image of microstructure of: a) as-cast, b) hot-forged, c) hot-rolled and d) QT state of S1100QL steel 3 RESULTS AND DISCUSSION 3.1 Temperature evolution during hot rolling In order to determine a semi-empirical equation for continuous cooling during the hot rolling of S1100QL, several laboratory rolled heats at the same re-heating temperature (1200 °C) and final thickness (12 mm) were required. No water de-scaling was used. Temperature evolution during the hot rolling of S1100QL can be cal- culated using the obtained semi-empirical equation: T = 1194.6 exp(–0.003t) (4) where T is the rolling temperature in °C and t is the time in seconds elapsed since the reheated billets were taken from reheating furnace. The temperature profile is shown in Table 2. From the rolling schedule, 7 roughing passes lie between 1158 °C and 971 °C followed by 4 finishing passes down to 882 °C. Achieved grains size and an example of the final as-rolled microstructure af- ter cooling from 882 °C in air is shown in Figure 1c and Figure 2. 3.2 Microstructure evolution Microstructures of as-cast, hot-forged, hot-rolled and heat-treated (QT) conditions are shown in Figure 1a to 1d, respectively. In the as-cast to hot-rolled condition, a mixed martensitic-bainitic microstructure can be ob- served. Martensite occurs predominantly in the segre- gated regions with solute enrichment. The mixed microstructure is related to relatively slow cooling rates achieved under natural cooling in air. The grain, except in the case of the as-rolled state, is equiaxial. In Fig- ure 1c the grains are elongated or "pancaked", indicating that the final rolling temperature (FRT) corresponds to the desired low FRT. The average PAGS and the distribu- tion of 400 grains evaluated for each condition are pre- sented in Figure 2. By combining the PAGS distribution for a separate process step, the average grain refinement from coarse 320 μm to rather fine 18 μm from as-cast to heat-treated condition is obvious. It is interesting to note that by con- tinuous hot rolling the maximum PAGS is still quite coarse in the range of 130 μm. Similar results were ob- tained after forging with a distribution factor of G avg /G max = 0.28 for rolled condition, where G stands for grain size in μm. Comparing the factor after re-austenitization with G avg /G max = 0.38, it becomes clear that re-austenitization of the forged-rolled condition has considerably improved the material isotropy by reducing the maximum grain size and narrowing the grain size distribution. This ex- plains why the toughness is sometimes improved by off-line heat treatment compared to on-line, with respect to the achieved austenite conditioning before on-line quenching. Nevertheless, isotropy is also improved when preferential grain orientation is changed from pan- cake-shaped to equiaxed by re-austenitization. The results of the PAGS for similar chemical compo- sitions but different processing routes, according to liter- ature, were from 30 μm to 7 μm. 4,29 PAGS in the hot-rolled condition is comparable to PAGS that Muckle- roy et al. 4 reported for the direct quenched, controlled rolled condition with a 40 % reduction below T nr , for steel of similar composition. They emphasised that martensitic blocks, as discussed before, determine the ef- fective grain size of martensitic steels and not PAGS, which is in accordance with earlier reports. 30,31 3.3 MFS analysis The results of the MFS analysis of laboratory hot-rolled S1100QL steel are presented in Figure 3. Rolling schedules and obtained MFS are shown in Ta- ble 2. Rolling consisted of 7 roughing passes (R1–R7), followed by 4 finishing passes (F1–F4). As can be seen in Figure 3, the MFS increases continuously up to R7 in a moderate manner, which is a result of a temperature drop and SRX. Limited secondary recrystallization or so-called grain growth appears in the case of R2, as esti- mated from the MFS curve. Rolling in the SRX region leads to grain refinement and improving the through-sec- J. FODER et al.: MEAN-FLOW-STRESS ANALYSIS OF LABORATORY HOT-ROLLED S1100QL STEEL ... Materiali in tehnologije / Materials and technology 54 (2020) 6, 901–908 905 Figure 3: MFS analysis for laboratory hot rolling of S1100QL steel. Regions of different recrystallization behaviour are marked in accor- dance with the literature 5,10 Figure 2: Prior austenite grain size distribution and average grain size for as-cast, hot-forged, hot-rolled and QT state of S1100QL steel tion grain size distribution by successive reductions with limited grain refinement. From R7 to F4 a distinct and almost linear MFS deflection is detected. The slope change was taken as the critical temperature defined as T nr and determined as an average between R7 to F1 due to the limited temperature raster taken by an individual test. If the limits are taken according to references 5,10 the MFS curve is interpreted as R1–R4 belong to Region I, which means that SRX refines PAGS during successive passes, given that the interpass time and deformation are sufficient. Nevertheless, according to the MFS curve, R1–R6 are taken as part of the roughing phase and F1–F4 is the finishing phase with R7–F1 being the pass in the region where Type III is expected and should be avoided. This could explain the rather coarse maximum PAGS found in the as-rolled state, knowing that during precipitation of Nb(C,N) niobium depletion appears in the matrix, affecting the grain-boundary mobility and en- hancing the recrystallisation, thus promoting grain growth. 32 If we apply temperatures calculated from equation (4), strain accumulation between R7 and F1 occurs at an average temperature of 958 °C. T nr , calculated according to the Boratto’s equation, is 936 °C, i.e., 22 °C lower than experimentally predicted, which seems reasonable. The results of the MFS analysis, i.e., the change in the slope after R7 to F4, can be associated with SEM mi- crographs of the hot-rolled condition, where elongated prior austenite grains can be seen. In the industrial practice, MFS analysis can easily be used to optimize rolling schedules based on the material response during hot rolling, even in real-case rolling pro- cesses. Due to the different starting conditions, such as slab geometry, size and reheating time and temperature, MFS values for industrial rolling are typically lower than the values determined during laboratory rolling. Also, RLT and RST, if assumed to be 150 MPa and 200 MPa, are considered as orientational values only, but according to the available data, this approximation still allows a fairly good estimation of where Region I and Region II are expected to be found. 5,10 Table 2: Per-pass reduction, calculated temperature, interpass time and MFS for laboratory hot rolling of S1100QL steel Pass no. Reduc- tion/% Tempera- ture/°C Interpass time/s MFS/MPa R1 5,0 1158 / 127 R2 11,4 1128 10 114 R3 13,6 1089 10 127 R4 15,8 1056 10 135 R5 17,4 1022 10 161 R6 16,8 993 9 174 R7 18,7 971 8 175 F1 17,1 945 8 231 F2 17,6 922 8 257 F3 14,3 903 8 309 F4 12,5 882 8 346 3.4 Thermodynamic analysis The predicted equilibrium phase diagram Fe-C has been simplified by excluding boron, sulphur and phos- phorous for easier visualisation. The minor addition of titanium shows no risk of primary (and coarse) TiN for- mation under the equilibrium condition of solidification due to the stability of nitrides below the solidus tempera- ture (presented as Liquid, see Figure 4) for the given chemical composition. The temperature stability of TiN in the carbon region of interest is practically unchanged due to the constant nitrogen content and covers the entire temperature region for austenite soaking and multi-pass hot-rolling for grain size control, including cooling after rolling. The solidification path for S1100QL is presented with a vertical line at 0.18 mass percent carbon. The phase field stability of TiN is presented by the prototype FCC_A1#3. The predicted isopleth diagram also shows temperature-carbon concentration relation of Nb(CN) precipitation based on the equilibrium thermodynamics of a multicomponent system according to the chemical composition in Table 1. The precipitation of Nb(C,N) in undeformed austen- ite is presented with the prototype of FCC_A1#2. With constant niobium and nitrogen concentration and in- creasing carbon, the temperature for the homogeneous nucleation of Nb(C,N) is increasing. This goes well with the nucleation theory, which refers to the solubility prod- ucts of niobium, carbon and nitrogen in austenite. Ac- cording to the prediction, the stability of Nb(C,N) with homogenic precipitation starts already under 1092 °C or 1104 °C using the TCFE7 or TCFE9 database, respec- tively. When using a thermodynamic approach for a multicomponent system, the temperatures are somewhat lower when calculated by the reference 25 , which only takes into account niobium, carbon and nitrogen, see Ta- ble 3. Taking into consideration the thermo-mechanical processing and dislocation density generated for poten- tial nucleation sites (based on precipitation kinetics and not included into the equilibrium phase stability calcula- J. FODER et al.: MEAN-FLOW-STRESS ANALYSIS OF LABORATORY HOT-ROLLED S1100QL STEEL ... 906 Materiali in tehnologije / Materials and technology 54 (2020) 6, 901–908 Figure 4: Isopletic equilibrium phase diagram for Fe-C for S1100QL tion) the total amount of precipitates at a given tempera- ture can vary significantly due to heterogenic nucleation. In this case it is also interesting to note that the experi- mentally determined T nr and the calculated one according to Boratto 22 are in a rather good agreement and both are lower than the predicted homogenic nucleation start for the precipitation of niobium-rich carbo-nitrides by Thermo-Calc. The determined starts of the homogeneous (equilibrium) and experimental precipitation of Nb(CN) are gathered in Table 3. The variation between the ther- modynamic calculations and the experimentally deter- mined T nr is expected, because the MFS change is related to the actual volume fraction and size of the precipitates formed by rolling, in order to effectively delay the recrystallisation kinetics and therefore to the rolling pa- rameters. Table 3: Calculated temperatures for Nb(C,N) precipitation and mea- sured T nr Thermo-Calc – Nb(C,N) precipitation TCFE7 1092,9 °C TCFE9 1104,4 °C Nb(C,N) precipitation according to Irvine 25 1122 °C T nr according to Boratto 22 936 °C MFS exp. – T nr 958 °C 3.5 Mechanical properties The mechanical properties of S1100QL steel in the QT condition are shown in Table 4. The yield strength after quenching and low-temperature tempering exceeds 1100 MPa, indicating that a pre-rolling strategy with ap- propriate pass numbers and re-heating temperatures be- fore hot rolling provides a good starting point for the QT process with off-line heat treatment. The ductility, ex- pressed in A 5,65 , also exceeds 10 %, which is required for this steel grade. Overall, the needed tensile properties are met. Charpy impact toughness KV 2 at –40 °C is 49 J, which is above 27 J, prescribed in the standard. 26 Table 4: Mechanical properties of S1100QL steel in QT state R p0,2 /MPa R m /MPa A 5,65 /% KV 2 /J min. 1100 1250–1500 min. 10 min. 27 1160 1377 13.5 49 3 CONCLUSIONS Laboratory hot rolling of S1100QL plate was suc- cessfully performed, and MFS analysis as an indirect method shows that the position of T nr is determined in the range of 150 MPa and 200 MPa. This agrees well for similar grades as the limitation for RLT and RST. A rela- tively good agreement was obtained by comparing the measured T nr and predicted T nr according to Boratto, with 958 °C and 936 °C, respectively. The tested rolling schedule reveals the presence of a locally coarse grain size and indicates that the optimization of rolling sched- ules is necessary for laboratory single stand rolling for the desired overall mechanical properties based on the obtained MFS curve presented in this paper. Significant improvements in the material isotropy were achieved by off-line heat treatment by reducing the maximum grain size and narrowing the grain size distribution. The data obtained with the given parameters can be used for rolling optimization with respect to the actual rolling mill capability in terms of the maximum allow- able force and torque as also for achieving the appropri- ate dimensional and flatness tolerance. Acknowledgments This research was made as a part of ^MRLJ research project co-financed by the Republic of Slovenia and the European Union under the European Regional Develop- ment Fund. The authors also want to acknowledge Douglas Stalheim, from DGS Metallurgical Solutions, the consul- tant of CBMM, for overall cooperation and the experi- ence exchange on the production of HSLA grades. 4 REFERENCES 1 Dillinger, Dillmax 1100 Datasheet. https://www.dillinger.de, 18.5. 2020 2 H. Mohrbacher, Property optimization in as-quenched martensitic steel by molybdenum and niobium alloying, Metals, 8 (2018)4 , doi:10.3390/met8040234 3 Y. Tomita, K. Okabayashi, Effect of microstructure on strength and toughness of heat-treated low alloy structural steels, Metallurgical and Materials Transactions A, 17 (1986) 7, 1203–1209, doi:10.1007/ BF02665319 4 N. C. Muckelroy, K. O. Findley, R. I. Bodnar, Microstructure and mechanical properties of direct quenched versus conventional reaustenitised and quenched plate, Journal of Materials Engineering and Performance, 22 (2013) 2, 512–522, doi:10.1007/s11665- 012-0251-y 5 D. G. Stalheim: Recrystallization behaviours in the production of structural steels, Proc. of the 52 nd Rolling Seminar, Rio de Jeneiro, 2015, 168–177, doi:10.5151/1983-4764-26355 6 Z. Wang, B. Wang, B. Wang, Y. Tian, T. Zhang, G. Yuan, Z. Liu, G. Wang, Development and application of thermo-mechanical control process involving ultra-fast cooling technology in China, ISIJ Inter- national, 59 (2019) 12, 2131–2141, doi:10.2355/isijinternational. ISIJINT-2019-041 7 D. Bomba~, M. Fazarinc, G. Kugler, S. Spaji}, Microstructure devel- opment of Nimonic 80A superalloy during hot deformation, Mate- rials and Geoenvironment, 55 (2008) 3, 319–328 8 A. Kri`aj, M. Fazarinc, M. Jenko, P. Fajfar, Hot workability of 95MnWCr5 tool steel, Materiali in tehnologije, 45 (2011)4 , 351–355 9 P. Fajfar, B. Brada{kja, B. Pirnar, M. Fazarinc, Determination of hot workability and processing maps for AISI 904L stainless steel, Mate- rials and Geoenvironment, 58 (2011) 4, 383–392 10 D. G. Stalheim, A. Gorni, M. M. Rebellato: Basic metallurgy/pro- cessing design concepts for optimized hot strip structural steel in yield strengths from 300 to 700 MPa, Proc. of the 53rd Rolling Semi- nar, Rio de Jeneiro, 2016, doi:10.5151/1983-4764-27546 11 B. Dutta, C. M. Sellars, Effect of composition and process variables on Nb(C, N) precipitation in niobium microalloyed austenite, Mate- rials Science and Technology, 3 (1987), 197–206, doi:10.1179/ 026708387790122846 J. FODER et al.: MEAN-FLOW-STRESS ANALYSIS OF LABORATORY HOT-ROLLED S1100QL STEEL ... Materiali in tehnologije / Materials and technology 54 (2020) 6, 901–908 907 12 S. F. Medina, A. Quispe, M. Gomez, Model of precipitation kinetics induced by strain for microalloyed steels, Steel research interna- tional, 76 (2005) 7, 527–531, doi:10.1002/srin.200506049 13 R. Sandström, R. Lagneborg, A model for hot working occurring by recrystallization, 23 (1975) 3, 387–398, doi:10.1016/0001-6160(75) 90132-7 14 S. Takebayashi, T. Kunieda, N. Yoshinaga, K. Ushioda, S. Ogata, Comparison of the dislocation density in martensitic steels evaluated by some x-ray diffraction methods, 50 (2010) 6, 875–882, doi:10.2355/isijinternational.50.875 15 L. P. Karjalainen, J. J. Jonas, Softening and flow stress behaviour of Nb microalloyed steels during hot rolling simulation, ISIJ Interna- tional, 35 (1995) 12, 1523–1531, doi:10.2355/isijinternational.35. 1523 16 H. Buken, E. Kozeschnik, A model for static recrystallization with simultaneous precipitation and solute drag, Metallurgical and Mate- rials Transactions A, 48A (2017), 2812–2818, doi:10.1007/s11661- 016-3524-5 17 J. Burja, M. Kole`nik, [. @uperl, G. Klan~nik, Nitrogen and nitride non-metallic inclusions in steel, Materiali in tehnologije, 53 (2019) 6, 919–928, doi:10.17222/mit.2019.247 18 L. Bäcke, Modeling of the effect of solute drag on recovery and recrystallization during hot deformation of Nb microalloyed steels, ISIJ International, 50 (2010) 2, 239–247, doi:10.2355/isijinter- national.50.239 19 S. Vervynckt, K. Verbeken, P. Thibaux, M. Thibaux, M. Liebeherr, Y. Houbaert, Austenite recrystallization-precipitation interaction in nio- bium microalloyed steels, SIJ International, 49 (2009) 6, 911–920, doi:10.2355/isijinternational.49.911 20 P. Gong, E. J. Palmiere, W. M. Rainforth, Characterisation of strain-induced precipitation behaviour in microalloyed steels during thermomechanical controlled processing, Materials Characterization, 124 (2017), 83–89, doi:10.1016/j.matchar.2016.12.009 21 A. Abdollah-Zedeh, D. P. Dunne, Effect of Nb on recrystallization after hot deformation in austenitic Fe–Ni–C, ISIJ International, 43 (2003) 8, 1213–1218, doi:10.2355/isijinternational.43.1213 22 R. Barbosa, F. Boratto, S. Yue, J. J. Jonas, The influence of chemical composition on the recrystallisation behaviour of microalloyed steels, Proceedings from an International Symposium on Processing, Microstructure and Properties of HSLA Steels, Pittsburgh, 1988, 51–61. 23 R. B. Sims, The calculation of roll force and torque in hot rolling mills, Proceedings of the Institution of Mechanical Engineers, 1954, 191–200, doi:10.1243/PIME_PROC_1954_168_023_02 24 F. Zhang, Y. Zhao, J. Shao, Rolling force prediction in heavy plate rolling based on uniform differential neural network, Journal of Con- trol Science and Engineering, (2016) 9, 1–9, doi:10.1155/2016/ 6473137 25 K. J. Irvine, F. B. Pickering, T. Gladman, Grain-refined C-Mn steels. Journal of Iron and Steel Institute, 205 (1967) 2, 161–182 26 EN 10025-6:2019 – Hot rolled products of structural steels – Part 6: Technical delivery conditions for flat products of high strength struc- tural steels in the quenched and tempered condition 27 EN ISO 6892-1:2020 – Metallic materials – Tensile testing – Part 1: Method of test at room temperature 28 EN ISO 148-1:2016 – Metallic materials – Charpy pendulum impact test – Part 1: Test method 29 A. Saastamoinen, A. Kaijalainen, D. Porter, P. Suikkanen, J. R. Yang, Y. T. Tsai, The effect of finish rolling temperature and tempering on the microstructure, mechanical properties and dislocation density of direct-quenched steel, Materials Characterization, 139 (2018), 1–10, doi:10.1016/j.matchar.2018.02.026 30 S. Morito, H. Yoshida, T. Maki, X. Huang, Effect of block size on the strength of lath martensite in low carbon steels, Materials Science and Engineering A, 438 (2006), 237–240, doi:10.1016/j.msea.2005. 12.048 31 J. W. Morris, Comments on the microstructure and properties of ultrafine grained steel, ISIJ International, 48 (2008), 8, 1063–1070, doi:10.2355/isijinternational.48.1063 32 B. Lopez, M. Rodriguez-Ibabe, Recrystallisation and grain growth in hot working of steels, Microstructure evolution in metal forming pro- cesses, Woodhead Publishing, (2012), 67–113 J. FODER et al.: MEAN-FLOW-STRESS ANALYSIS OF LABORATORY HOT-ROLLED S1100QL STEEL ... 908 Materiali in tehnologije / Materials and technology 54 (2020) 6, 901–908