Next Article in Journal
Efficient Roll-Forming Simulation Using Non-Conformal Meshes with Hanging Nodes Handled by Lagrange Multipliers
Previous Article in Journal
Effect of Glue, Thiourea, and Chloride on the Electrochemical Reduction in CuSO4–H2SO4 Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Temperature Oxidation of Steel under Linear Flow Rates of Air and Water Vapor—An Experimental Determined Set of Data

Department of Ferrous Metallurgy, Montanuniversitaet Leoben, Franz-Josef-Straße 18, 8700 Leoben, Austria
*
Author to whom correspondence should be addressed.
Metals 2023, 13(5), 892; https://doi.org/10.3390/met13050892
Submission received: 5 April 2023 / Revised: 27 April 2023 / Accepted: 3 May 2023 / Published: 5 May 2023
(This article belongs to the Section Corrosion and Protection)

Abstract

:
High-temperature oxidation phenomena play a crucial role in steel production and processing. The development of new alloying concepts, the modification of production routes due to the decarbonization of steelmaking, and increasing quality demands will additionally stimulate research activities on high-temperature oxidation. Within the scope of this publication, the oxidation behavior of an unalloyed steel was investigated by means of thermogravimetric (TG) and metallographic analysis. The TG measurements were performed with a simultaneous thermal analyzer in combination with a water vapor generator and a mass spectrometer for quality control. Experiments were conducted in a temperature range from 900 to 1200 °C under various oxidation atmospheres, containing synthetic air and water vapor, with different linear flow rates of the oxidizing medium. According to the thermogravimetric data, an evaluation method is introduced to obtain direct kinetic data for linear and parabolic scale growth. The influence of the experimental parameters on high-temperature oxidation is discussed in detail.

1. Introduction

Over recent decades, high-temperature oxidation of iron and steel has been a widely studied topic to establish an improved knowledge of the underlying mechanisms [1,2,3,4,5]. During steel production and processing, the product surface is exposed to temperatures up to 1250 °C and a wide variation of oxidizing atmospheres. Generally, high-temperature oxidation phenomena may be divided into external and internal. Various internal oxidation mechanisms can be observed, especially in alloyed steels or steels contaminated by trace elements, which may lead to essential quality losses of the product [6,7,8,9,10,11,12,13,14]. Current trends in the steel industry, such as the increased use of obsolete scrap, the corresponding increase of the content of residuals and tramp elements, or the replacement of natural gas with hydrogen for industrial burners, just to name a few, will raise new oxidation issues. The present paper introduces an experimental setup, designed to investigate the influence of widely varying thermal cycles and atmospheres on high-temperature oxidation. Combining the gas supply, water vapor generator, simultaneous thermal analysis, and mass spectrometer provides the precise adjustment of experimental conditions and delivers highly reproducible measurement results. A wide variety of industrial processes for steel production may thus be simulated in the future, including continuous casting, reheating, hot rolling, and annealing.
The first Investigations mainly address external oxidation, the calculation of growth rates for linear and parabolic scale growth, the influence of temperature on scale formation, and the impact of increasing water vapor contents on growth rates. The results also indicate that experimental parameters, such as gas flow rate, significantly affect high-temperature oxidation rates and, therefore, need to be adjusted carefully. The present work provides an excellent basis for extending future research activities toward simulating internal oxidation phenomena, such as grain boundary oxidation or liquid metal embrittlement.

1.1. Oxidation Kinetics and Influencing Factors

The progression of scale layer thickness, “x”, can be described in a simplified way by the linear and parabolic rate law in Equations (1) and (2). In 1 and 2, kl denotes the linear and kp the parabolic rate constant, t is the time in seconds, and Cl and Cp are constant values. Tammann was the first to derive the parabolic growth law from Fick’s first law in 1920 [15,16]. For applications to industrial processes, oxidation is often described similarly by referring to weight gain per area, “W.” Pilling and Bedworth initially described the parabolic relationship in Equation (3) in 1923 [17]. The integration constants are omitted under the boundary condition that the scale layer thickness or the mass gain are zero at time t = 0. Table 1 summarizes the SI units of x, W, kl, and kp. Values for kl and kp are usually determined via thermogravimetric methods, where changes in mass due to the reaction of the base material with oxygen are recorded [18,19,20,21,22,23,24,25]. The layer and mass-related scaling constants can be converted into each other.
d x d t = k l x   x = k l x t + C l x
d x d t   ~   J = D   Δ c x   o r   d x d t = k p x x   x 2 = 2 k p x t + C p x
d W d t = k p W W   W 2 = 2 k p W t + C p W
A linear, uniform increase in scale layer thickness can be observed mainly at the beginning of the oxidation. Only a thin oxide layer is present, through which iron ions are transported quickly. The scaling rate is controlled by the transfer and adsorption/dissociation of oxygen from the oxidation medium to the scale surface and the oxidation reaction itself. As the oxidation time becomes longer, the scale layer thickness increases, which increases the transport path of iron ions and subsequently reduces the quantity available for the oxidation reaction. As a result, the diffusion-controlled transport of iron ions for the oxidation reaction becomes the limiting factor, and the scaling follows a parabolic rate law [23,24,26,27,28].
The oxidation rate is influenced significantly by the chemical composition of steel, especially by the elements chromium and silicon [19,29,30,31,32,33,34,35,36,37,38,39], and experimental parameters, such as temperature [5,19,21,22,23,24,25,40,41,42,43], surface condition [44], type of gas, and gas quantity. The influence of water vapor and the gas flow rates are explained in more detail.

1.1.1. Water Vapor

In steel processing, water vapor is part of most oxidizing atmospheres, such as in continuous casting resulting from the air–mist–cooling of the strand surface, in reheating furnaces from combustion, and in rolling from cooling water, just to name a few examples. The presence of water vapor in an oxidation atmosphere significantly increases the scaling rate [21,38,41,45,46,47,48,49]. A detailed explanation of the underlying mechanism was performed by Rahmel and Tobolski [45]. As the scale grows, the plasticity decreases and local fissures form at the interface between the metal and oxide due to flow obstructions, similar to the “edge effect” reported by Engell [50]. A supporting effect of the colliding oxide layers occurs at the corners, which impedes the flow of the scale layer and causes it to detach from the metal surface. Thus, the gap between the metal surface and the scale layer inhibits further diffusion of iron ions and decreases the oxidation rate. Water vapor can penetrate the gaps formed in this way and react with the metal surface to metal oxide and hydrogen, which diffuses to the oxide layer and reduces it. The resulting water vapor can diffuse back to the metal surface while the iron ions and electrons migrate outwards through the oxide layer. This process repeats until oxide bridges, which span the resulting gaps, are formed. The resulting pores in the oxide layer show that the water vapor/hydrogen mixture has degraded through reduction. Tuck et al. [47] found that the formation of pores in the scale does not have to occur in an oxidation atmosphere containing water vapor, but scaling is nevertheless increased compared to pure oxygen. The increased plasticity of the scale explains this behavior due to a better movement of dislocations, preventing an early loss of contact between the scale and steel.

1.1.2. Flow Rate

The amount of the oxidation medium supplied per unit of time is crucial for the oxidation progress. As already mentioned, iron diffuses fast during the phase of linear oxidation. A higher gas velocity at the interface accelerates the transport from the gas medium to the scale, the adsorption and dissociation, and finally, the oxidation rate. If the quantity of oxidation medium provided at the interface is equal to or greater than the quantity of iron ions transported by diffusion, the oxidation speed becomes independent of the further increased quantity and velocity of the oxidation gas. This is referred to as a critical gas velocity [24,28].

2. Materials and Methods

2.1. Equipment and Methods

All thermogravimetric oxidation experiments were conducted using a Netzsch STA 449 F3 Jupiter (STA, Simultaneous Thermal Analysis, Netzsch GmbH & Co. Holding KG, Selb, Germany) with a DTA-TG (Differential thermal analysis—Thermogravimetric analysis) sample carrier for freely suspended samples. The complete setup includes various precision pressure controllers and mass flow controllers (MFCs, Bronkhorst High-Tech B.V., Ruurlo, The Netherlands), a water vapor generator aSteam DV-2 (aDROP Feuchtemeßtechnik GmbH, Fürth, Germany), a quadrupole mass spectrometer QMS 403 C Aëolos (Netzsch GmbH & Co. Holding KG, Selb, Germany), several thermostats (Netzsch GmbH & Co. Holding KG, Selb, Germany), and various control valves (Netzsch GmbH & Co. Holding KG, Selb, Germany). Figure 1 shows a schematic flow chart of the gases and the system components and Figure 2 is a schematic illustration of the experimental setup. The gases arrive first at the top side of the sample and a shielding gas (argon) is supplied on the bottom of the furnace, which is permanently switched on to protect the balance system. The balance system has an accuracy of 0.1 µg and continuously records the mass change during experiments. Temperature control is performed by two type S thermocouples located in the sample holder and beside the sample with an accuracy of +/−2 °C.
At the gas outlet, a partial gas flow is continuously analyzed by the mass spectrometer, so the adjusted gas atmosphere can be controlled; see Figure 3. Only impeccable experiments are evaluated regarding the gas supply, and misinterpretations due to a delayed gas can be avoided, as shown in Figure 3b, where the water vapor was delayed for approximately 5 min.
The amount of gas supplied by the MFCs is up to 500 mL·min1 and their linear precision and repeatability precision is ±1% F.S. (percentage of Full Scale) and ±0.2% F.S., respectively. Contrary to the gases provided by the MFCs, water vapor is directly supplied from the water vapor generator. The operating principle can be summarized as follows: distilled water from the integrated water tank is fed into an electrically heated evaporator unit. The following chamber mixes the resulting water vapor with an added carrier gas. An additional heater ensures a preheating of the carrier gas and that the temperature in the mixing chamber does not fall below the dew point of the resulting gas mixture. Without a carrier gas, it is possible to perform oxidation experiments with 100% saturated steam. The maximum gas amount of H2O corresponds to a value of approximately 415 mL·min−1. Thermostats with a temperature of 200 °C are used at the transfer lines to avoid condensation in the system after the water vapor generator. The uniformity of the water vapor flow rate is controlled by the mass spectrometer, so constancy can be ensured over a complete experiment; see Figure 3.
In the experimental case, an isobaric system can be assumed, whereby Gay-Lussac’s law can be applied to calculate the supplied gas volume at each oxidation temperature under the assumption that thermal expansion for each oxidation gas is equal. For a given gas volume, the gas velocity in the furnace tube (inner diameter 26.5 mm) can be calculated from the cross-sectional area reduced by the space requirement of the sample and sample holder. For simplicity, a homogeneous speed distribution is assumed in the furnace. The flow rate of shielding gas, which arrives from the balance system, is not considered in those calculations because it never reaches the sample.

2.2. Oxidation Program and Parameters

In the framework of the research presented, pure (technical) iron was investigated. The chemical composition is shown in Table 2.
This raw material was cut with the wet abrasive cutting machine ATM Brilliant 220 (ATM Qness GmbH, Mammelzen, Germany) to prepare samples measuring 13 mm × 12.2 mm × 2 mm. To suspend the specimen, a hole measuring 2.5 mm in diameter was drilled in the upper part of the sample. No additional grinding or polishing took place after cutting and the surface roughness corresponded to a Sa value of approximately 1 µm. In contrast to the Ra value, the Sa value represents the (3D) areal average roughness. For subsequent examinations, the exact dimensions of each sample were measured. To avoid contamination of the sample, a cleaning step with ethanol and acetone was performed right before each experiment. Preceding an experiment, the furnace chamber, balance system, and gas supply lines were evacuated three times and purged with argon. This ensured identical initial conditions for each test and eliminates the risk of pre-oxidation during the inert heating process. Prior to the experiment, a baseline compensation with the same time–temperature–gas program was performed with an inert Al2O3 sample to eliminate the influence of the incident gas flow and the thermal buoyancy on the weight signal.
Figure 4 shows the time–temperature–gas program for the oxidation tests performed. All samples were slowly heated to the target temperature (900−1000−1100−1200 °C) at 15 °C·min1, followed by a homogenization for 5 min, ensuring uniform conditions at the sample surface. Heating and homogenization were performed under argon 5.0 purging at a constant purging rate. Immediately afterwards, the oxidation gas, either synthetic air (80 vol.-%N2 and 20 vol.-%O2) or a synthetic air/water vapor mixture, was injected into the furnace and the oxidation was started for a period of 60 min. The parameters for the oxidation experiments are summarized in Table 3. After the oxidation time, the gas supply was switched back to inert to preserve the state of high-temperature oxidation and avoid post-oxidation. The furnace tube was flushed simultaneously with argon 5.0 and nitrogen, and the sample was cooled with a cooling rate of −60 °C·min1.
The structure of the final scale layer was investigated by metallographic methods. The samples were cold embedded before cutting to ensure that the scale and steel/scale interface was protected. After the cutting operation, another cold embedding was carried out for easier handling during grinding (grit 180−320−600−1200) and polishing (9 µm and 3 µm). The samples were analyzed using an optical light microscope (Polyvar Reichert-Jung MEF 2 and Keyence VHX7000, KEYENCE International, Mechelen, Belgium).

3. Results and Discussion

Table 3 summarizes all relevant parameters for the oxidation experiments. Unless otherwise stated, the flow rate refers to the calibration temperature of the MFC. Depending on the gas quantity and oxidation temperature, the gas velocity in the furnace tube ranged from 0.15 to 9.35 cm·s−1. The Reynolds number was determined with respect to the temperature and gas velocity. A maximum value of approximately 8 indicates a completely laminar gas flow; therefore, laminar conditions can be assumed for all experiments.
The first step to a uniform evaluation of all experimental results is to define the onset of the mass gain curve, determined by the intersection of two tangents [51]. The first tangent is applied in the time range of the inert atmosphere and corresponds to a horizontal straight line. The second tangent is applied in the initial mass gain range, whereby smaller discontinuities of the measurement curve are not considered. The defined onset point represents the starting point for the standardized generation of mass gain curves. A mathematical description based on the results of the oxidation experiments is performed by using a regression model. As mentioned, the scaling of steel is described by a combination of linear and parabolic oxidation.
To obtain the linear oxidation constant, kl, the initial part of the measured mass gain curve is fitted by linear regression (Equation (4)) using the data points of the first 100 s. If a transition from linear to parabolic oxidation takes place before 100 s, the range of the measuring points considered must be adjusted. Afterwards, the time, tt, where a transition to diffusion-controlled growth occurs can be determined precisely by analyzing the deviation squares from the mass gain curve and the linear regression at each measuring point. The time that exceeds a deviation higher than a pre-defined factor represents the start of diffusion-controlled oxidation. The pre-defined boundary factor is usually set to 0.1, but if required, it is adjusted individually to achieve the best-fitting results. To compensate discontinuities in the measurement curve, especially in the initial phase, the deviation squares are averaged over a range of 25 s.
W = k l t
The mathematical description of the diffusion-controlled part is performed via a parabolic function. Due to dividing the regression curve into two areas, an additional coefficient “a” is needed to ensure a continuous function; see Equation (5). The slopes of the linear and parabolic oxidation regime must be identical at the transition time, tt, and it can be described mathematically by the first derivative of the two functions at this point; see Equation (6). By using a starting value for kp, a first value for a can be determined. The exact evaluation of kp and a is performed iteratively by minimizing the deviation squares between the measured mass gain curve and the parabolic regression. The evaluation is entirely automatic and usually does not require any manual correction. Thus, uncertainties that would arise from a human evaluation are avoidable. The complete evaluation procedure is shown schematically in Figure 5.
W 2 = k p t + a   W = k p t + a
k l = k p 2 k p t t + a a = k p 2 k l 2 k p t t
Figure 6a shows the results of the weight increase for the test series from 900 to 1200 °C. The dashed lines for the oxidation temperatures 900 and 1200 °C indicate an experiment repetition to ensure reproducibility. For better comparability, the mass gain is related to the sample surface. The flow rate of synthetic air is 200 mL·min−1. Based on the investigated steel grade and the experimental temperatures, a scale that mainly consists of wustite is expected but in the area of the scale/gas interface, magnetite and hematite may also form [5,52,53]. An increase in temperature leads to an enhancement of oxidation due to increased diffusion. The mass increase for 900 °C is approximately 28 mg·cm−2 after 60 min. For 1000 °C, it is 49 mg·cm−2, and for 1100 °C, 81 mg·cm−2. The mass gain at 1200 °C is approximately 110 mg·cm−2. The slope of the linear oxidation segment is only slightly influenced by the temperature, indicating that the transfer from the gas and adsorption/dissociation of oxygen is also a rather fast process at 900 °C. However, a lower temperature leads to a significantly faster transition from linear to parabolic oxidation due to the limited diffusion of iron ions. The scale thickness quickly reaches a level at which diffusion is not sufficient for a constant mass gain. The values obtained for kl, kp, and tt are summarized in Table 4 and Figure 6b.
In order to ensure that the results are independent of the gas velocity in the furnace tube, the purging rate of air was varied for 900 and 1200 °C. The results for 1200 °C are shown in Figure 7a,b. In addition to the 200 mL·min−1 purging rate, as commonly used (3.74 cm·s−1 at 1200 °C), tests were carried out for 25, 100, 400, and 500 mL·min−1. This corresponds to a gas velocity in the furnace tube of 0.47, 1.87, 7.48, and 9.35 cm·s−1, respectively.
For the lowest gas velocity, the linear and parabolic oxidation segments are significantly influenced by the purging rate, whereby a linear mass gain is observed for almost the whole experiment. The total mass gain is approximately 80 mg·cm−2. An increase in gas velocity intensifies oxidation drastically in the early oxidation stage, and an earlier transition from a linear growth characteristic to a parabolic one occurs. The transition time, tt, ranges from 84 to 2468 s and increases significantly with decreasing gas velocity. In the parabolic range, oxidation is hardly influenced by gas velocity and the parabolic rate constant stays approximately the same. Thus, the mass gain curves equalize, and the final mass gain does not depend on the gas purging rate. Consequently, if the oxidation time exceeds 30 min, a gas velocity in the furnace tube of around 1.87 cm·s−1 is sufficient to achieve results irrespective of the gas quantity, resulting in a final mass gain of roughly 110 mg·cm−2 after 60 min. In the linear oxidation segment, the differences in the mass gain become smaller for a gas velocity higher than 7.48 cm·s−1. For the experiments at 1200 °C with a gas velocity higher than 3.74 cm·s−1, a discontinuity in the mass gain curve becomes visible after about 2000 s. At this time, the “edge effect” due to flow obstruction of the scale causes the scale layer on one side to lift off completely. The oxidation rate decreases later during the experiment because of the interrupted iron ion diffusion path, as the mass gain on this site is limited to the oxygen uptake during wustite transformation to magnetite and hematite.
Figure 8a,b show the mass gain at 900 °C for a gas velocity from 0.15 up to 5.96 cm·s−1 and correspond to a flow rate of 10 up to 400 mL·min−1. As in the experiments at 1200 °C, an increase in gas velocity intensifies oxidation, especially in the early (linear) stage. The mass gain is notably influenced at the lowest investigated gas velocity of 0.15 cm·s−1. For a gas velocity higher than 0.37 cm·s−1, a similar mass gain is achieved after 60 min and the parabolic rate constant stays approximately the same. The final mass gain is around 28 mg·cm−2. In the linear range, the differences in the mass gain get smaller for a gas velocity higher than 4.47 cm·s−1. The transition time, tt, decreases from 1247 s for the lowest gas velocity to 19 s for the highest gas velocity. Table 5 and Figure 9 summarize the values for the oxidation constants, kl and kp, and the transition time, tt.
According to the literature section, the results of the gas variation experiments at 900 and 1200 °C can be explained as follows (see Table 5 and Figure 9). A higher gas velocity at the sample surface favors mass transfer and adsorption/dissociation of oxygen. Thus, a high gas velocity results in initially faster oxidation (higher kl). However, oxidation cannot be increased to infinity because areas where adsorption/dissociation can take place are limited, so linear oxidation becomes more similar for higher gas velocities; see Figure 7 and Figure 8. Related to kl, the transition time, tt, at high gas velocities is far less affected compared to at low gas velocities. The necessary mass gain for a transition from linear to parabolic oxidation increases slightly with decreasing gas velocity due to the fact that the amount of iron ions required for linear growth can be maintained longer since mass transfer and adsorption/dissociation of oxygen are lower; see Figure 7 and Figure 8. During parabolic oxidation, the diffusion of iron ions controls the growth rate of the oxide layer. The thicker the layer, the slower the growth rate and vice versa. Hence, the different oxide layer thicknesses balance after a while and kp is approximately steady. This occurs when the gas supply is higher than the critical gas velocity. As a consequence of the lower diffusion at 900 °C, kl, kp, and tt are shifted to lower values compared to 1200 °C. However, the trends of the determined data remain the same; see Figure 9.
In a study by Abuluwefa et al. [24], at a temperature of 1150 °C, a critical gas velocity of 4.3 cm·s−1 was determined for an overall constant mass gain in the parabolic range. Below this value, the gas velocity influenced the oxidation rate significantly. Compared to the value obtained by Abuluwefa et al. [24], the values in this study are significantly smaller. In studies from the early 1930s, values for the critical velocity at 927 and 1230 °C were determined with 2.7 and 2.0 cm·s−1, respectively [5,54,55]. The obtained value of 1.87 cm·s−1 at 1200 °C is similar to this value, but at 900 °C, it (0.37 cm·s−1) differs significantly. According to the results, the range of the gas velocity influencing parabolic oxidation drops with decreasing oxidation temperature because of the lower diffusivity of the iron ions.
For oxidation experiments in laboratories (e.g., thermogravimetric analysis), the critical gas velocity is an essential parameter for the investigation of oxidation experiments as the gas supply can significantly influence the results, especially for short oxidation times. For industrial processes, the gas velocity is difficult to describe respectively since, in many cases, neither the atmosphere, nor the speed of oxidation medium supply, can be controlled entirely. Furthermore, turbulent atmospheric conditions (e.g., secondary cooling zone in continuous casting) will probably lead to location-dependent high gas velocities. Thus, the data generated in laboratory experiments improve the overall understanding of the process and may serve as input data for industrial process simulations.
Figure 10a shows the resulting mass gain for water vapor variation experiments at 1200 °C. The purging rate of synthetic air is 200 mL·min−1 for all experiments. A higher volume fraction of water vapor means a higher total gas quantity. The experiment with 50.1 vol.-%water vapor corresponds to a theoretical gas quantity of 548 mL·min−1 at 100 °C and was performed twice. The rate constant for linear oxidation is almost independent of the water vapor content. Only in atmospheres that contain higher water vapor contents (above 30.7 vol.%), a slight increase of kl noticeable. It can be concluded that adsorption/dissociation of O2 on the scale surface is the time-controlling step for kl and the dissociation of H2O is considerably slower due to more complex reactions at the interface. However, the transition from linear to parabolic oxidation is smoother in its presence. The linear/parabolic transition time, tt, is in the range of 120 to 173 s. Compared with the previously explained experimental series, there is no clear pattern for the transition time, tt. For up to 40.3 vol.-%water vapor in the oxidation medium, tt is almost unaffected and basically corresponds to that of pure synthetic air. The slight variations are due to measurement errors. The presence of water vapor favors parabolic scaling and consequently improves the diffusion of the iron ions through the scale. However, in the range of linear scaling, this is not yet decisive since the provision of oxygen is the rate-determining step. As a result, there is no noticeable change in the transition time. For the highest water vapor content in the oxidation medium, there is a slight decrease in the transition time, which cannot be explained with the hypothesis, just discussed. The meager increase in kl should not be responsible for the decrease of tt. Consequently, it can be concluded that a further mechanism is affecting the transition time, which has to be studied in more detail. For a water vapor content of 9.4 vol.%, the mass gain increases to approximately 145 mg·cm−2, which corresponds to an increase of approximately 30% compared to synthetic air. A further rise in the water vapor content to 20.6 vol.-%results in a mass gain of 161 mg·cm−2. At a water vapor content of 30.7, 40.3, and 50.1 vol.%, the mass gain increases to 170, 175, and 180 mg·cm−2, respectively. The increased scaling in water vapor atmospheres is attributable to the mechanisms described in the literature section. With increasing water vapor content in a mixture with synthetic air, the influence of water on the mass gain decreases. Scale detachments are compensated in the presence of water vapor resulting in a larger area for diffusion of iron ions to the outside and, therefore, in a massive increase in scaling. The differences for lower water vapor contents in the oxidation medium are particularly distinct since the compensation is just starting. With further increasing water vapor content in the oxidation medium, the scale mostly remains in complete contact with the specimen surface during oxidation. As a result, the outward diffusion of iron ions remains almost constant and the influence of water vapor on scaling decreases at higher contents. Table 6 and Figure 10b summarize the values attained for kl, kp, and tt.
Figure 11a,b shows a comparison of the steel/scale interface for the oxidation temperatures of 900 and 1200 °C and 200 mL·min−1 synthetic air after 3600 s of oxidation. The scale layer at 900 °C is still in contact with the steel surface, while at 1200 °C, a lift-off from the steel surface is notable. Due to the presence of oxygen and the ongoing reactions at the outer surface of the scale, wustite is transformed to magnetite and finally to hematite. The transformation is visible as a bright gray layer on the scale outside; see Figure 11b. The optical interpretation of the wustite transformation is based on an evaluation performed by Chen et al. [46]. The scale layer itself hardly shows any pores at 900 °C, while the complete scale layer at 1200 °C is covered with pores. For the 1200 °C sample, an incorrect metallographic preparation can largely be excluded since the other examined samples do not show any accumulation of pores. In addition, a localized occurrence of the pores in the scale is recognizable, which also contradicts the metallographic preparation as the source. A higher tendency to pore formation in the detached scale was found among others by Engell et al. [50]. The change in the scale morphology could be explained by the progressing transformation of wustite and the necessary diffusion processes of the iron ions to reach the reaction interfaces. Furthermore, density changes and changes in the crystal structure of the scale layer occur. However, it should be noted that Chen et al. [46] found no increased tendency to the formation of porosities in a detached scale during the wustite transformation. Compared to the experimental parameters in this publication, the experiments performed by Chen et al. [46] were at lower temperatures, which could be a reason for the differences. The exact processes leading to this occurrence of the scale cannot be clearly determined based on this series of experiments alone and further investigations are still required. The scales at 1000 and 1100 °C are similar to the 900 °C sample.
The samples oxidized in a mixture of air and water vapor, shown in Figure 12a,b, show a significantly denser scale with a lower quantity of pores than those that oxidized in synthetic air. The appearing pores are located at the steel/scale interface. It can be concluded that the void formation theory described by Rahmel and Tobolski [45] is not primarily responsible for the increased oxidation in the presence of water vapor for the present experiments. As mentioned by Tuck [47], the appearance of voids in the scale does not always occur. Nevertheless, oxidation is still increased in water vapor due to the increased flowability of the scale, which ensures a permanent contact of the steel and scale, resulting in a higher outward diffusion of iron ions.

4. Conclusions

In the framework, pure iron was isothermally oxidized in a temperature range from 900 to 1200 °C in various atmospheres containing synthetic air and water vapor and with different linear flow rates. The following results can be highlighted:
  • An iterative regression model was presented to attain values for kl, kp, and the transition time tt where oxidation changes from linear to parabolic. The measured mass gain curve is approximated by a linear and parabolic function, considering that the slope of both functions is identical at the transition point linear/parabolic to ensure an overall continuous function.
  • For oxidation with synthetic air, an increase in temperature only leads to a slight increase in the linear oxidation constant, kl, whereas the transition time, tt, increases significantly. Furthermore, the parabolic oxidation constant, kp, becomes considerably larger with increasing temperature.
  • A higher water vapor content in a constant flow rate of synthetic air significantly increases scaling respectively the parabolic oxidation constant, kp. The linear oxidation constant, kl, is almost independent of the water vapor content, indicating that the linear oxidation is mainly controlled by the oxygen from synthetic air.
  • In some cases, the values determined for the critical gas velocity differ significantly from those from the literature. For synthetic air at 1200 °C, an increase in gas velocity above 1.87 cm·s−1 does not result in a higher mass gain after 60 min. The parabolic rate constant, kp, stays approximately the same. At 900 °C, a value of 0.37 cm·s−1 is determined as the critical gas velocity. A higher gas velocity in the linear oxidation segment significantly influences scaling behavior, especially for lower gas velocities.

Author Contributions

Conceptualization, P.P., G.G. and C.B.; Methodology, G.G. and P.P.; Resources, G.G.; Writing—Original Draft Preparation, G.G.; Visualization, G.G.; Investigation, G.G.; Writing—Review & Editing, G.G., P.P. and C.B.; Supervision, C.B. and P.P.; Funding Acquisition P.P. and C.B.; Project Administration, P.P. and C.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the COMET program “Integrated Computational Material, Process and Product Engineering (IC-MPPE)” Project No 859480. The APC was funded by the Department of Ferrous Metallurgy, Montanuniversitaet Leoben.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study is available on request from the corresponding author.

Acknowledgments

The authors gratefully acknowledge the assistance of the Austrian Federal Ministries for Climate Action, Environment, Energy, Mobility, Innovation and Technology (BMK) and for Digital and Economic Affairs (BMDW), represented by the Austrian research funding association (FFG), and the federal states of Styria, Upper Austria, and Tyrol.

Conflicts of Interest

The authors declare no conflict of interest.

Correction Statement

This article has been republished with a minor correction to resolve typographical errors. This change does not affect the scientific content of the article.

References

  1. Birks, N.; Meier, G.H.; Pettit, F.S. Introduction to the High-Temperature Oxidation of Metals, 2nd ed.; Cambridge University Press: Cambridge, UK, 2006. [Google Scholar]
  2. Gao, W.; Li, Z. Developments in High-Temperature Corrosion and Protection of Materials; Woodhead Publishing Limited: Cambridge, UK, 2008. [Google Scholar]
  3. Young, D.J. High Temperature Oxidation and Corrosion of Metals, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2016. [Google Scholar]
  4. Sequeira, C.A.C. High Temperature Corrosion: Fundamentals and Engineering; Wiley: Hoboken, NJ, USA, 2019. [Google Scholar]
  5. Chen, R.Y.; Yuen, W.Y.D. Review of the High-Temperature Oxidation of Iron and Carbon Steels in Air or Oxygen. Oxid. Met. 2003, 59, 433–468. [Google Scholar] [CrossRef]
  6. Bernhard, C.; Linzer, B.; Presoly, P.; Watzinger, I.; Watzinger, J. Evaluation of AHSS concepts with a focus on the product properties and appropriate casting characteristics of Arvedi ESP thin slab casters. In Proceedings of the 5th International Conference on Advances in Solidification Processes, Salzburg, Austria, 17–21 June 2019. [Google Scholar]
  7. Liu, Y.; Yang, C.; Chai, F.; Pan, T.; Su, H. High Temperature Oxidation Resistance of 9Ni Steel. J. Iron Steel Res. Int. 2014, 21, 956–963. [Google Scholar] [CrossRef]
  8. Zeze, M.; Tanaka, A.; Tsujino, R. Formation Mechanism of Sliver-type Surface Defect with Oxide Scale on Sheet and Coil. Tetsu-To-Hagane 2001, 87, 85–92. [Google Scholar] [CrossRef][Green Version]
  9. Höfler, T.; Auinger, M.; Linder, B.; Angeli, G.; Danninger, H. Analysing and Modeling High Temperature Oxidation of Steel Slabs during Reheating. In Proceedings of the 11th International Rolling Conference, São Paulo, Brazil, 1–3 October 2019. [Google Scholar]
  10. Trindade, V.B.; Krupp, U.; Wagenhuber, P.E.-G.; Christ, H.-J. Oxidation mechanisms of Cr-containing steels and Ni-base alloys at high-temperatures -. Part I: The different role of alloy grain boundaries. Mater. Corros. 2005, 56, 785–790. [Google Scholar] [CrossRef]
  11. Yin, L.; Sridhar, S. Effects of Small Additions of Tin on High-Temperature Oxidation of Fe-Cu-Sn Alloys for Surface Hot Shortness. Metall. Mater. Trans. B 2010, 41, 1095–1107. [Google Scholar] [CrossRef]
  12. Osei, R.; Lekakh, S.; O’Malley, R. Effect of Cu Additions on Scale Structure and Descaling Efficiency of Low C Steel Reheated in a Combustion Gas Atmosphere. Oxid. Met. 2022, 98, 363–383. [Google Scholar] [CrossRef]
  13. Yeşiltepe, S.; Şeşen, M.K. High-temperature oxidation kinetics of Cu bearing carbon steel. SN Appl. Sci. 2020, 2, 643. [Google Scholar] [CrossRef][Green Version]
  14. Mikl, G.; Höfler, T.; Gierl-Mayer, C.; Danninger, H.; Linder, B.; Angeli, G. Scaling Behaviour of Si-alloyed Steel Slabs under Reheating Conditions. JCME 2021, 5, 71–74. [Google Scholar] [CrossRef]
  15. Tammann, G. Über Anlauffarben von Metallen. Z. Anorg. Allgem. Chem. 1920, 111, 78. [Google Scholar] [CrossRef]
  16. Bürgel, R.; Maier, H.-J.; Niendorf, T. Handbuch Hochtemperatur-Werkstofftechnik: Grundlagen, Werkstoffbeanspruchungen, Hochtemperaturlegierungen und -Beschichtungen; Vieweg+Teubner: Wiesbaden, Germany, 2011. [Google Scholar]
  17. Pilling, N.; Bedworth, R.J. The Oxidation of Metals at High Temperatures. Inst. Met. 1923, 29, 529. [Google Scholar]
  18. Nilsonthi, T. Oxidation behaviour of hot-rolled Si-containing steel in water vapour between 600 and 900 °C. In Proceedings of the 10th Thailand International Metallurgy Conference, Bangkok, Thailand, 10–13 July 2018. [Google Scholar]
  19. Alaoui Mouayd, A.; Koltsov, A.; Sutter, E.; Tribollet, B. Effect of silicon content in steel and oxidation temperature on scale growth and morphology. Mater. Chem. Phys. 2014, 143, 996–1004. [Google Scholar] [CrossRef][Green Version]
  20. Othman, N.K.; Zhang, J.; Young, D.J. Water Vapour Effects on Fe–Cr Alloy Oxidation. Oxid. Met. 2010, 73, 337–352. [Google Scholar] [CrossRef]
  21. Yun, J.-Y.; Ha, S.-A.; Kang, C.-Y.; Wang, J.-P. Oxidation Behavior of Low Carbon Steel at Elevated Temperature in Oxygen and Water Vapor. Steel Res. Int. 2013, 84, 1252–1257. [Google Scholar] [CrossRef]
  22. Wikström, P.; Blasiak, W.; Du, S.C. A study on oxide scale formation of low carbon steel using thermo gravimetric technique. Ironmak. Steelmak. 2008, 35, 621–632. [Google Scholar] [CrossRef]
  23. Abuluwefa, H.T.; Guthrie, R.I.L.; Ajersch, F. Oxidation of low carbon steel in multicomponent gases: Part II. Reaction mechanisms during reheating. Metall. Mater. Trans. A 1997, 28, 1643–1651. [Google Scholar] [CrossRef]
  24. Abuluwefa, H.; Guthrie, R.I.L.; Ajersch, F. The effect of oxygen concentration on the oxidation of low-carbon steel in the temperature range 1000 to 1250 °C. Oxid. Met. 1996, 46, 423–440. [Google Scholar] [CrossRef]
  25. Aghaeian, S.; Brouwer, J.C.; Sloof, W.G.; Mol, J.M.; Böttger, A.J. Numerical Model for Short-Time High-Temperature Isothermal Oxidation of Fe–Mn Binaries at High Oxygen Partial Pressure. High Temp. Corros. Mater. 2023, 99, 201–218. [Google Scholar] [CrossRef]
  26. Pettit, F.; Wagner, J. Transition from the linear to the parabolic rate law during the oxidation of iron to wüstite in CO-CO2 mixtures. Acta Metall. 1964, 12, 35–40. [Google Scholar] [CrossRef]
  27. Smeltzer, W. The kinetics of wustite scale formation on iron. Acta Metall. 1960, 8, 377–383. [Google Scholar] [CrossRef]
  28. Rahmel, A. Die Oxydation von Eisen bei hohen Temperaturen in Inertgasen mit kleinen Sauerstoffgehalten. Mater. Corros. 1972, 23, 95–98. [Google Scholar] [CrossRef]
  29. Yang, Y.-L.; Yang, C.-H.; Lin, S.-N.; Chen, C.-H.; Tsai, W.-T. Effects of Si and its content on the scale formation on hot-rolled steel strips. Mater. Chem. Phys. 2008, 112, 566–571. [Google Scholar] [CrossRef]
  30. Adachi, T.; Meier, G.H. Oxidation of iron-silicon alloys. Oxid. Met. 1987, 27, 347–366. [Google Scholar] [CrossRef]
  31. Hao, M.; Sun, B.; Wang, H. High-Temperature Oxidation Behavior of Fe-1Cr-0.2Si Steel. Materials 2020, 13, 509. [Google Scholar] [CrossRef] [PubMed][Green Version]
  32. Kahveci, A.I.; Welsch, G.E. Oxidation of Fe-3 wt.% Cr alloy. Oxid. Met. 1986, 26, 213–230. [Google Scholar] [CrossRef]
  33. Li, Z.-F.; Cao, G.; He, Y.; Liu, Z.; Wang, G.-D. Effect of Chromium and Water Vapor of Low Carbon Steel on Oxidation Behavior at 1050 °C. Steel Res. Int. 2016, 87, 1469–1477. [Google Scholar] [CrossRef]
  34. Liu, X.; Cao, G.; He, Y.; Jia, T.; Liu, Z. Effect of Temperature on Scale Morphology of Fe-1.5Si Alloy. J. Iron Steel Res. Int. 2013, 20, 73–78. [Google Scholar] [CrossRef]
  35. Suarez, L.; Schneider, J.; Houbaert, Y. High-Temperature Oxidation of Fe- Si Alloys in the Temperature Range 900–1250 °C. Defect Diffus. Forum 2008, 273, 661–666. [Google Scholar]
  36. Mehtani, H.K.; Khan, M.I.; Jaya, B.N.; Parida, S.; Prasad, M.; Samajdar, I. The oxidation behavior of iron-chromium alloys: The defining role of substrate chemistry on kinetics, microstructure and mechanical properties of the oxide scale. J. Alloy. Compd. 2021, 871, 159583. [Google Scholar] [CrossRef]
  37. Gutierrez-Platas, J.L.; Artigas, A.; Monsalve, A.; García-Gómez, N.A.; García-Rincón, O.; De la Garza-Garza, M.; Pérez-González, F.A.; Colás, R.; Garza-Montes-De-Oca, N.F. High-Temperature Oxidation and Pickling Behaviour of HSLA Steels. Oxid. Met. 2018, 89, 33–48. [Google Scholar] [CrossRef]
  38. Chen, J.; Shen, Z.; Zhang, J. Effect of Water Vapor and Oxygen Partial Pressure on Oxidation of Fe and Fe–Cr Alloys. High Temp. Corros. Mater. 2023, 99, 15–46. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Zhang, C.; Li, F.; Wang, Z.; Wang, X.; Wang, C.; Huang, J.; Mao, F.; Chen, C.; Jiang, T.; et al. High-Temperature Oxidation Behavior of Cr-Ni-Mo Hot-Work Die Steels. Materials 2022, 15, 5145. [Google Scholar] [CrossRef]
  40. Zhang, S.; Long, M.; Zhang, H.; Songyuan, A.; Chen, D.; Liu, P.; Duan, H. Experimental study on evolution of oxidation behavior on high-temperature continuous casting slab surface. J. Mater. Res. Technol. 2021, 11, 2049–2058. [Google Scholar] [CrossRef]
  41. Chen, R.Y.; Yuen, W.Y.D. Effects of the Presence of Water Vapour on the Oxidation Behaviour of Low Carbon–Low Silicon Steel in 1%O2–N2 at 1173 and 1273 K. Oxid. Met. 2013, 79, 679–699. [Google Scholar] [CrossRef]
  42. Zhang, H.; Yu, L.; Liu, T.; Ni, H.; Li, Y.; Chen, Z.; Yang, Y. Optimizing the preheating temperature of hot rolled slab from the perspective of the oxidation kinetic. J. Mater. Res. Technol. 2020, 9, 12501–12511. [Google Scholar] [CrossRef]
  43. Arreola-Villa, S.A.; Vergara-Hernández, H.J.; Solorio-Diáz, G.; Pérez-Alvarado, A.; Vázquez-Gómez, O.; Chávez-Campos, G.M. Kinetic Study of Oxide Growth at High Temperature in Low Carbon Steel. Metals 2022, 12, 147. [Google Scholar] [CrossRef]
  44. Pineda Huitron, R.M.; Ramírez López, P.E.; Vuorinen, E.; Nazen Jalali, P.; Pelcastre, L.; Kärkkäinen, M. Scale Formation on HSLA Steel during Continuous Casting Part II: The Effect of Surface Conditions. Metals 2020, 10, 1245. [Google Scholar] [CrossRef]
  45. Rahmel, A.; Tobolski, J. Einfluss von Wasserdampf und Kohlendioxyd auf die Oxydation von Eisen in Sauerstoff bei hohen Temperaturen. Corros. Sci. 1965, 5, 333–346. [Google Scholar] [CrossRef]
  46. Chen, R.Y.; Yuen, W.Y.D. Effects of the Presence of Water Vapour on the Oxidation Behaviour of Low Carbon–Low Silicon Steel in 1%O2–N2 at 1073 K. Oxid. Met. 2013, 79, 655–678. [Google Scholar] [CrossRef]
  47. Tuck, C.W.; Odgers, M.; Sachs, K. The oxidation of iron at 950 °C in oxygen/water vapour mixtures. Corros. Sci. 1969, 9, 271–285. [Google Scholar] [CrossRef]
  48. Chen, Y.R.; Liu, Y.; Li, C. Comparison of the oxidation behaviour of 60Si2MnA at 800–1200 °C in dry and wet air. Mater. High Temp. 2020, 37, 279–294. [Google Scholar] [CrossRef]
  49. Pineda Huitron, R.M.; Ramírez López, P.E.; Vuorinen, E.; Nazen Jalali, P.; Pelcastre, L.; Kärkkäinen, M. Scale Formation on HSLA Steel during Continuous Casting Part I: The Effect of Temperature–Time on Oxidation Kinetics. Metals 2020, 10, 1243. [Google Scholar] [CrossRef]
  50. Engell, H.; Wever, F. Über einige Grundfragen der Bildung und der Haftung von Zunder auf Eisen. Acta Metall. 1957, 5, 695–702. [Google Scholar] [CrossRef]
  51. Hemminger, W.F.; Cammenga, H.K. Methoden der Thermischen Analyse; Springer: Berlin, Germany, 1989. [Google Scholar]
  52. Davies, M.H.; Simnad, M.T.; Birchenall, C.E. On the Mechanism and Kinetics of the Scaling of Iron. J. Met. 1951, 3, 889–896. [Google Scholar] [CrossRef]
  53. Païdassi, J. Sur la précipitation de magnetite aux limites des joints de la sous-structure d’un protoxyde de fer riche en oxygène. Acta Metall. 1958, 6, 69–71. [Google Scholar] [CrossRef]
  54. Jominy, W.E.; Murphy, D.W. Scaling of steel at forging temperatures. Trans. Amer. Soc. Steel Treat. 1930, 18, 19. [Google Scholar]
  55. Upthegrove, C. Engineering Research Bulletin; No. 25; George Banta Publication: Menasha, WI, USA, 1933. [Google Scholar]
Figure 1. Flow chart of the experimental setup.
Figure 1. Flow chart of the experimental setup.
Metals 13 00892 g001
Figure 2. Schematic illustration of the furnace.
Figure 2. Schematic illustration of the furnace.
Metals 13 00892 g002
Figure 3. Example of a recorded mass spectrometer signal during the oxidation. (a) Oxidation medium is on time at the start of oxidation. (b) Oxidation medium is delayed at the start of oxidation.
Figure 3. Example of a recorded mass spectrometer signal during the oxidation. (a) Oxidation medium is on time at the start of oxidation. (b) Oxidation medium is delayed at the start of oxidation.
Metals 13 00892 g003
Figure 4. Investigated time–temperature–gas program.
Figure 4. Investigated time–temperature–gas program.
Metals 13 00892 g004
Figure 5. Schematic illustration for the generation of kl, tt, and kp. (a) Linear fit of the initial part of the mass gain curve. (b) Analyzing the deviation squares between the measured mass gain curve and the linear fit to determine tt. (c) Iterative fit to identify the most appropriate parabolic function.
Figure 5. Schematic illustration for the generation of kl, tt, and kp. (a) Linear fit of the initial part of the mass gain curve. (b) Analyzing the deviation squares between the measured mass gain curve and the linear fit to determine tt. (c) Iterative fit to identify the most appropriate parabolic function.
Metals 13 00892 g005
Figure 6. (a) Comparison of the mass gain at different oxidation temperatures for 200 mL·min−1 synthetic air. (b) Visualization of kl, kp, and tt.
Figure 6. (a) Comparison of the mass gain at different oxidation temperatures for 200 mL·min−1 synthetic air. (b) Visualization of kl, kp, and tt.
Metals 13 00892 g006
Figure 7. (a) Comparison of the mass gain in synthetic air for different gas velocities at 1200 °C. (b) Detail of the linear oxidation.
Figure 7. (a) Comparison of the mass gain in synthetic air for different gas velocities at 1200 °C. (b) Detail of the linear oxidation.
Metals 13 00892 g007
Figure 8. (a) Comparison of the mass gain in synthetic air for different gas velocities at 900 °C. (b) Detail of the linear oxidation.
Figure 8. (a) Comparison of the mass gain in synthetic air for different gas velocities at 900 °C. (b) Detail of the linear oxidation.
Metals 13 00892 g008
Figure 9. Visualization of kl, kp, and tt obtained at different gas velocities for (a) 900 °C and (b) 1200 °C.
Figure 9. Visualization of kl, kp, and tt obtained at different gas velocities for (a) 900 °C and (b) 1200 °C.
Metals 13 00892 g009
Figure 10. (a) Comparison of the mass gain for a mixture of 200 mL·min1 synthetic air with different water vapor contents. (b) Visualization of kl, kp, and tt.
Figure 10. (a) Comparison of the mass gain for a mixture of 200 mL·min1 synthetic air with different water vapor contents. (b) Visualization of kl, kp, and tt.
Metals 13 00892 g010
Figure 11. (a) Synthetic air at 900 °C. (b) Synthetic air at 1200 °C.
Figure 11. (a) Synthetic air at 900 °C. (b) Synthetic air at 1200 °C.
Metals 13 00892 g011
Figure 12. (a) Synthetic air + 9.4 vol.-%water vapor at 1200 °C. (b) Synthetic air + 30.6 vol.-%water vapor at 1200 °C.
Figure 12. (a) Synthetic air + 9.4 vol.-%water vapor at 1200 °C. (b) Synthetic air + 30.6 vol.-%water vapor at 1200 °C.
Metals 13 00892 g012
Table 1. SI units of x, W, kl, and kp.
Table 1. SI units of x, W, kl, and kp.
x k l x k p x W k p W
mm·s−1m2·s−1kg·m−2kg2·m−4·s−1
Table 2. Chemical composition analyzed via optical emission spectroscopy.
Table 2. Chemical composition analyzed via optical emission spectroscopy.
C [wt.%]Mn [wt.%]Si [wt.%]P [wt.%]S [wt.%]
0.01270.0167<0.0010.00450.0038
Table 3. Experimental parameters.
Table 3. Experimental parameters.
ExperimentTemperature [°C]AtmosphereTime [min]Flow Rate
[mL·min−1]
Gas Velocity b
[cm·s−1]
1900synthetic air60100.15
2900synthetic air60250.37
3900synthetic air601001.49
4 (I; II)900synthetic air602002.98
5900synthetic air603004.47
6900synthetic air604005.96
71000synthetic air602003.23
81100synthetic air602003.48
91200synthetic air60250.47
101200synthetic air601001.87
11 (I, II)1200synthetic air602003.74
121200synthetic air604007.48
131200synthetic air605009.35
141200synthetic air with 9.4 vol.-%water vapor60302 a4.13
151200synthetic air with 20.6 vol.-%water vapor60344 a4.71
161200synthetic air with 30.6 vol.-%water vapor60392 a5.37
171200synthetic air with 40.3 vol.-%water vapor60458 a6.27
18(I, II)1200synthetic air with 50.1 vol.-%water vapor60548 a7.51
a flow rate at 100 °C; b gas velocity at oxidation temperature; (I; II) experiment performed twice.
Table 4. Values for kl, kp, and tt in 200 mL·min1 synthetic air for different temperatures.
Table 4. Values for kl, kp, and tt in 200 mL·min1 synthetic air for different temperatures.
Temperature
[°C]
Gas Mediumkl
[mg·cm−2·s−1]
tt
[s]
kp
[mg2·cm−4·s−1]
900synthetic air0.1229 0.17
10000.15850.43
11000.161261.10
12000.171692.56
Table 5. Values for kl, kp, and tt at 900 and 1200 °C for different gas velocities.
Table 5. Values for kl, kp, and tt at 900 and 1200 °C for different gas velocities.
900 °C
Gas velocity [cm·s−1]0.150.371.49 2.98 4.47 5.96
kl [mg·cm−2·s−1]0.0090.020.080.120.150.17
tt [s]124733554292319
kp [mg2·cm−4·s−1]0.050.170.160.170.170.18
1200 °C
Gas velocity [cm·s−1]0.471.87 3.74 7.489.35
kl [mg·cm−2·s−1]0.030.100.170.260.31
tt [s]24683101699684
kp [mg2·cm−4·s−1]1.322.442.562.482.41
Table 6. Values for kl, kp, and tt at 1200 °C in a mixture of 200 mL·min1 synthetic air with different water vapor contents.
Table 6. Values for kl, kp, and tt at 1200 °C in a mixture of 200 mL·min1 synthetic air with different water vapor contents.
Temperature
[°C]
Gas Mediumkl
[mg·cm−2·s−1]
tt
[s]
kp
[mg2·cm−4·s−1]
1200synthetic air with 0.0 vol.-%H2O0.171692.56
with 9.4 vol.-%H2O0.171573.68
with 20.6 vol.-%H2O0.171736.80
with 30.6 vol.-%H2O0.181588.08
with 40.3 vol.-%H2O0.181728.64
with 50.1 vol.-%H2O0.1912010.44
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gaiser, G.; Presoly, P.; Bernhard, C. High-Temperature Oxidation of Steel under Linear Flow Rates of Air and Water Vapor—An Experimental Determined Set of Data. Metals 2023, 13, 892. https://doi.org/10.3390/met13050892

AMA Style

Gaiser G, Presoly P, Bernhard C. High-Temperature Oxidation of Steel under Linear Flow Rates of Air and Water Vapor—An Experimental Determined Set of Data. Metals. 2023; 13(5):892. https://doi.org/10.3390/met13050892

Chicago/Turabian Style

Gaiser, Georg, Peter Presoly, and Christian Bernhard. 2023. "High-Temperature Oxidation of Steel under Linear Flow Rates of Air and Water Vapor—An Experimental Determined Set of Data" Metals 13, no. 5: 892. https://doi.org/10.3390/met13050892

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop