You are on page 1of 12

AISTech 2019 — Proceedings of the Iron & Steel Technology Conference

6–9 May 2019, Pittsburgh, Pa., USA


DOI 10.1000.377.271

Analysis of the Effects of Oxygen Enrichment in a Reheating Furnace

Bethany M. Worl1, Jianglin Fan1, Armin Silaen1, Jeffrey Cox2, Kurt Johnson3, Larry Fabina2, Joe Maiolo4, Kelly Tian4,
Chenn Q. Zhou1
1
Center for Innovation through Visualization and Simulation (CIVS), Purdue University Northwest
2200 169th Street, Hammond, IN 46323, USA
Phone: 1-219-989-2665
Email:czhou@pnw.edu
2
ArcelorMittal-USA
250 W US Hwy 12, Burns Harbor, IN 46304, USA
3
ArcelorMittal Global Research and Development
3001 E Columbus Dr, East Chicago, IN 46312, USA
4
Praxair, Inc.
175 E Park Dr, Tonawanda, NY 14150, USA

Keywords: Reheat furnace, computational fluid dynamics, CFD, oxygen enrichment, air/fuel ratio

INTRODUCTION
In 2017, the iron and steel industries in the U.S produced 23 million metric tons of pig iron and 82 million tons of steel; the
total value of iron and steel produced was $147 billion [1]. In the steelmaking process, the reheat furnace consumes the largest
amount of energy in the hot rolling stage and the second highest amount of energy in the entire process [2]. This energy
consumption is necessary for the heating of steel products up to the temperatures desired for plastic deformation in the hot
rolling stage of around 1150°C or greater [3]. The steel must also obtain a uniform temperature distribution to produce high
quality rolled steel and avoid damage to the rollers [4]. In order to reach these high temperatures and reduce the temperature
gradient, steel slabs may reside within a reheating furnace for over two hours. Reheated steel products can be in the form of
billets, blooms or slabs. There are many different avenues of optimization due to the myriad operational parameters found
within the reheating furnace.
The main objectives of reheating furnace optimization are to maintain satisfactory product quality, increase furnace
efficiency, increase productivity, and reduce emissions. The efficiency of the reheating furnace is commonly defined as the
amount of energy that enters the steel divided by the total furnace heat input. The efficiency can vary according to production
and delays, but is typically between 45-60%, with around 35-53% of the combustion energy released going to the slabs [5-8].
Product quality is quantified by the temperature distribution within the reheated slab but can also be adversely affected by
decarburization and oxidation within the furnace [9]. A temperature distribution within the steel can also damage the rollers
due to the differing plasticity within the slab. Within the steelmaking process, the reheat furnace is second to the blast furnace
in terms of greenhouse gas emissions, which include CO2, CO, and NOx, among other trace pollutants [5]. CO2 is a natural by-
product of the combustion process for hydrocarbons, whereas NOx is caused by excess oxygen reacting with nitrogen at the
very high temperatures found near the combustion flame. Production is also a value that many seek to optimize so long as
product quality does not suffer.
Improvement of thermal efficiency is often sought after as it should decrease emissions and increase productivity by nature of
its definition. The largest loss of heat from the furnace, and thus the largest detractor of the efficiency, is from the flue gases.
Most reheating furnace operations salvage energy from the flue gas using a recuperator to preheat the air being used in
combustion. Oxy-fuel and oxygen enriched combustion have also shown to be a somewhat novel method of reducing energy
loss by reducing the volume of the flue gas and the removal, in part or entirely, of nitrogen from the combustion reaction.
Reduction of nitrogen, typically 70% of the volume of outgoing flue gases, reduces possible NOx formation and leads to

© 2019 by the Association for Iron & Steel Technology. 2629


easier processing of other pollutants from their increased concentration [10-11]. This increased volume of H2O and CO2 in the
furnace gases also enhances heat transfer to the slab as both species participate in radiative heat transfer and nitrogen does
not. As radiation contributes over 90% of the heat transfer to the slab, thermal efficiency is greatly improved [12]. Karimi and
Saidi studied the heat transfer and energy efficiency for a pusher-type reheat furnace using oxygen enhanced air for
combustion [13]. Notable energy consumption savings and NOx reduction were found from an increase in volumetric oxygen
content from 21% to 60%. Han et al. recently found that oxy-fuel combustion for coke oven gas was 54% more efficient than
air-fuel combustion for an average increase in efficiency of 27.6% [14].
As computing technology and infrastructure has advanced, the utilization of computational fluid dynamics techniques (CFD)
to simulate real world phenomena has become quite popular. The main advantage of CFD is the ability to numerically work
on problems in a process without the material waste and downtime previously required for troubleshooting. Some
experimentation is required in order to validate the numerical model, but this experimentation is much less wasteful and
potentially more elucidatory than exhaustive trials could prove to be. As a common, high-temperature industrial process,
simulation of the reheating furnace has been undertaken for decades. Kim reviewed the main engineering models that have
been developed [15]. Of the three given, CFD modeling is the most complex and expansive, though this accuracy does come at
the cost of computation time.
Prieler et al. have recently well predicted the heating characteristics of steel products with a novel iterative method between a
steady-state gas phase combustion simulation with a transient billet heating process [16-17]. Of particular interest, then, are
their later studies into oxy-fuel and scale formation modeling. Mayr et al. modeled billets in a pusher-type reheating furnace
as a highly viscous fluid [18]. With the addition of oxy-fuel burners and oxygen enrichment, productivity was found to
increase from 60 t/h to 80 t/h with a further increase in efficiency from 62.9% to 65%. Schluckner et al. developed a local and
time-dependent scale formation model that utilizes experimental data on 2 types of steel in an air-fuel, oxygen enriched, and
oxy-fuel environments [19]. As expected, an increase in temperature and concentration of oxidizing species in the flue gas
(CO2, H2O, O2) caused by oxygen enrichment and oxy-fuel combustion resulted in increased specific mass gain of scale
when all trials ran the same amount of time. However, the efficiency increase resulting from oxygen enriched combustion
allowed the steel products to reach the final dropout temperature much faster and so scale formation was reduced due to
much reduced furnace residence time. Previously scale formation had only been modeled using CFD by Jang et al. who
assumed a purely parabolic kinetic consideration in their scale formation model [20].
CFD models for the reheating furnace and for scale formation have both been developed. These models may now be used to
illuminate common industrial problems and to predict performance under certain improvements. This paper reports on the
modeling of the complex combustion and heat transfer phenomena within a reheating furnace under different conditions as
they relate to air/fuel ratio and oxygen enrichment. Additionally, it reports on the application of a previously proposed scale
formation model used to predict and analyze the effects that operational changes may have on yield due to material loss [21]. A
case is made for controlled scaling used to provide benefits in surface quality in order to purge casting slivers or defects.

METHODOLOGY AND MODELING


The numerical simulation presented in this paper follows the general procedure of geometry creation, meshing, CFD
simulation setup and solving, and post-processing. The numerical calculation is a process of using iterative methods to solve
the discretized Navier-Stokes and scalar equations through high performance computers. All simulations were run to
convergence as determined by temperature points, mass flow rate, and residuals.

Numerical models
Simulations were run using 3-dimensional, double precision, pressure based solvers for both the steady and transient cases
using ANSYS Fluent® CFD software. Mass and momentum conservation through their respective equations was ensured
through convergence criteria. The turbulence was modeled using the Reynolds Averaged Navier-Stokes (RANS) realizable k-
ε model. The energy equation was enabled to include heat transfer in the model. Necessary for combustion, the species
transport model with the eddy dissipation concept (EDC) model was used. The EDC model gives detailed chemical reactions
within the predicted flows. As radiation constitutes over 90% of the heat transfer into the slab, a discrete ordinates (DO)
radiation model was used in conjunction with the Weighted-Sum-of-Gray-Gases model (WSGGM) to capture the gray gas
behavior.

Scale formation models


The scale build-up and heat transfer models were explored in a previous conference paper and publication [21]. For brevity, a
quick review follows.

2630 © 2019 by the Association for Iron & Steel Technology.


Two main kinetic phenomena, gas-phase transport and solid-state diffusion, contribute to high-temperature scale formation
like that found within the industrial reheating furnace. Gas-phase transport, resulting from the reaction of oxidizing species
with the bare iron or steel, was approximated mathematically as a linear function [22]. Solid-state diffusion, the process of iron
and oxygen diffusing through the layers of iron oxide found in a mill scale, occurs once scale has been built up and was
mathematically approximated as a parabolic function [22]. Modeled as serial resistances, both processes were incorporated into
a single mathematical function [23]:

𝑥 𝑥 𝑡 (1)

where t is time, 𝑘 is the linear rate constant, and 𝑘 is the parabolic rate constant. The x term is either scale thickness or
weight gain per unit area, depending on the derivation of the rate constants and their subsequent units. The discretized form
of this equation is quickly found:

1
2
1
x ∑ ∙ ∆𝑡 (2)
1 2 4𝑡
𝑘𝑙 𝑘𝑝

where all terms retain their value from before, ∆𝑡 is the time step of the simulation, and t is the total time elapsed.
The heat transfer model for the scale formation made use of a modified thermal conductivity of the outermost layer of steel so
as to reflect the combined resistances of both the steel and the growing oxide layer. This was accomplished through an
effective thermal conductivity given by:

𝑘 (3)

where L is the thickness of either the steel or the scale (subscript s) and k is the thermal conductivity of the same. To be clear,
𝐿 is the thickness of the numerical cell being modified with the effective thermal conductivity by a user-defined function
(UDF) and is not the entire slab thickness. As this function takes the insulating effects of the iron oxide into effect, there is no
need to geometrically model the scale within the simulation.
Both of the scale formation model parts are implemented through the use of a UDF within ANSYS Fluent.

GEOMETRY AND BOUNDARY CONDITIONS


In this study, a 3-zone pusher-type reheating furnace was modeled. The model of the reheating furnace was validated with
industrial data in previous publications using both gas temperature and oxygen volume fraction [24]. The furnace was divided
into zones, namely: preheating, heating, and soaking zones. Figure 1 details the location of all burners, doors, outlets,
supporting structures, slabs, and zone locations. Due to the nature of the pusher-type reheating furnace, there were no gaps
between slabs. Slabs are nonuniform but have average dimensions of 9.9” x 55” x 318”. The furnace modeled was roughly
112 ft long and 34 ft wide and had 70 natural gas and air burners across all 3 zones. In order to improve the efficiency of
numerical calculation, the gas tubes and some other structures connected with the burners were not included in the model.

Figure 1.Computational domain of the pusher-type reheat furnace.


Due to the size and complexity of the furnace geometry, about 13 million elements were necessary to achieve acceptable
accuracy. Mesh density was highest in the main cyclone of each burner where fuel mixes with air. Furnace walls were
assumed adiabatic with an emissivity of 0.75 due to the negligible losses found in real-world production [25]. Table 1 details

© 2019 by the Association for Iron & Steel Technology. 2631


the fuel and air inputs per zone (top and bottom) for the baseline case. These inputs were for the entirety of the burners found
in the reported section (i.e. the fuel mass flow rate reported for the preheat zone top is the summation of the fuel inputs for all
10 burners in that zone section) and are hourly-averaged fuel and air flow rates during a certain operation period. As such,
validation for this model was also an hourly-average number to be consistent. For transient simulations, the slab movement
was modeled using a UDF and a time-varying velocity to account for the different slab resident times. Heat losses caused by
skids were also modeled through a location-based UDF based on industrial temperature data.

Table 1. Fuel and air boundary conditions

Primary fuel Secondary fuel Fuel Primary air Secondary air Air
mass flow rate mass flow rate temperature mass flow rate mass flow rate temperature
(ft3/s) (ft3/s) (°F) (ft3/s) (ft3/s) (°F)
Preheat zone
33.088 1.741 333.904 83.476
top

Preheat zone
26.514 1.395 271.792 67.948
bottom

Heat zone
13.936 0.733 131.424 32.856
top
80 811.3
Heat zone
9.756 0.513 92.096 23.024
bottom
Soak zone
2.251 0.118 23.112 5.778
small

Soak zone
4.370 0.230 41.232 10.308
large

Two different reaction mechanisms for the natural gas combustion were used in this simulation. The two-step Westbrook and
Dryer (WD) reaction mechanism was used for the study of air/fuel ratio, as it has been found to be an accurate assessment of
air-fuel flames [26]. Unfortunately, this accuracy is lacking when used to model oxygen enriched and oxy-fuel combustion.
Yin et al. found that the refined Jones and Lindstedt (JL) 4-step reaction model was better able to predict hydrogen levels,
flame temperature, and the carbon monoxide level in oxygen enriched combustion [27]. As such, the latter reaction mechanism
was used to investigate oxygen enrichment and oxy-fuel combustion undertaken in the latter half of the study. Both
mechanisms are summarized in Table 2 and in further detail in Ref. 27.

Table 2. Reaction mechanism for fuel gas

Pre-exponential Activation Energy E


Reactions
Factor A (1/s) (J/mol)

Westbrook and Dryer 2-step CH4 + 3/2 O2 → CO + 2 H2O 5.012 x 1011 2.0 x 108
mechanism CO + 1/2 O2 → CO2 2.239 x 1012 1.7 x 108
CH4 + 1/2 O2 → CO + 2 H2 4.4 x 1011 1.26 x 108

Refined Jones and Lindstedt CH4 + H2O → CO + 3H2 3.0 x 108 1.26 x 108
4-step mechanism H2 + 1/2 O2 ↔ H2O 5.69 x 1011 1.465 x 108
CO + H2O ↔ CO2 + H2 2.75 x 109 8.36 x 107

Scale formed once temperatures exceed 1292°F (700°C) has been found to be around 95% wüstite (FeO) [28]. As such, the
mathematical relationship for the thermal conductivity of wüstite above 550°C proposed by Akiyama et al. was used in the
calculation for the effective thermal conductivity [29].

2632 © 2019 by the Association for Iron & Steel Technology.


RESULTS
Air/fuel Ratio
The composition of the product gases change with increased oxygen enrichment of combustion air. As previously mentioned,
this alters the radiative heat transfer occurring within the furnace due to increased concentration of radiating species and the
reduction of non-radiating species such as nitrogen. Nitrogen also does not react during the combustion reaction, and so much
of the heat from combustion goes to bring the nitrogen up to the same temperature as the products of the combustion reaction.
A preliminary air/fuel ratio study was done to study the effect of nitrogen on the combustion mechanism and product quality
when differing conditions were imposed. This entailed changing the air flow rates while keeping the fuel flow rates constant
in all 3 zones to study fuel rich and fuel lean environments as indicated in Table 3.

Table 3. Case matrix of AFR study

Average Air/fuel Ratio Average Oxygen/fuel Ratio


Case
(mass basis) (molar basis)
1 16 1.86
Stoichiometric 17.6 2.05
2 18 2.10
3 20 2.33

4 22 2.56

The air/fuel ratio is an important factor with respect to combustion as it affects furnace gas temperature and resulting
emissions. Figure 2 details the change in the bulk temperature of the slab throughout the entire furnace residence time of each
case. In progressive order, the boundaries of the preheating, heating, and soak zones are denoted by the dashed vertical lines
at 76 minutes and 117 minutes. The model predicts a negligible increase in the bulk slab temperature profile with increasing
AFR until the slab approaches the heating zone. The largest temperature difference was found at the end of the furnace
residence time and was most likely due to changes in the gas temperature in the soak zone. Case 1 has the lowest slab bulk
temperature of 2287°F, which was expected for the fuel-rich substoichiometric combustion ratio. Fuel, entering the furnace
at ambient temperatures is heated to the furnace gas temperature but remains uncombusted and so does not provide heating
energy to the slab. Conversely, the AFR of case 2 that is closest to stoichiometric gives the highest slab bulk temperature of
2352°F. Increased air fraction results in lower gas temperatures as unreacted oxygen and higher nitrogen ballast dilute the
energy of combustion.

2500
Preheating Heating Soak
Slab Bulk Temperature (°F)

2000

AFR: 16
1500
2400 AFR: 18
2350 AFR: 20
1000
2300
AFR: 22
2250
500
2200
140 145 150 155
0
0 50 100 150 200
Residence Time (min)

Figure 2. Slab bulk temperature through furnace residence for AFR cases

© 2019 by the Association for Iron & Steel Technology. 2633


The AFR potentially also impacts product quality and furnace throughput due to scale formation. As AFR increases, the
concentration of free oxygen within the furnace invariably increases. As oxygen is more oxidizing than water vapor or carbon
dioxide, the amount of oxidation and subsequent scale formation is expected to increase with increasing AFR. Figure 3 shows
said increase in predicted average scale thickness with increasing AFR; levels of excess air from 0-20% are shown on the
figure. Obviously, the deleterious effects of scale formation on steel must be included when analyzing changes in furnace
combustion. Also shown is an increase and then decrease in slab bulk temperature with AFR peaking just past stoichiometric.
The effect on slab temperature has been discussed. Excess air is commonly used in industrial burners in order to completely
burn off all of the fuel to avoid pollutants such as carbon monoxide the cooling effect of unburned fuel [30]. The increased
mixing the excess air induces was the primarily prescribed reason for the slab temperature to be highest just above the ideal
stoichiometric combustion scenario. This slab top temperature would mirror the flame temperature through the resulting
radiative heat into slab. The turbulence and combustion models used along with their underlying assumptions could have also
had an effect on the results.

0.08 2360

2350
0.07
Final scale thickness (in)

Slab bulk temperature (°F)


2340

0.06 2330

2320
0.05 2310

20% Excess Air 
10% Excess Air 
0% Excess Air 
(AFR = 17.6)

(AFR = 20.6)
(AFR = 18.9)

2300
0.04
2290

0.03 2280
14 16 18 20 22 24
Air/fuel ratio
Figure 3. Comparison of slab top temperature and final scale thickness with AFR
The combustion profile is also affected by a change in the AFR. Using CO concentration as an indicator of the extent of
combustion provides insight into the mixing of oxygen and fuel. Figure 4 shows the decreasing trend in the maximum CO
concentration near the burner with increasing AFR. This information was taken from a plane 2 inches from the burner face
where mixing between the fuel and air was underway. At larger AFR the extent of combustion increased as indicated by the
decrease in CO concentration. The lowest extent of combustion is at the highest CO concentration corresponding to the
substoichiometric AFR. The decrease in CO concentration between AFR 16 and 18 is at first steep but then begins to lessen
with each increment upwards of the AFR. Each upward increment of air/fuel ratio (mass basis) shown (i.e. 16  18) is
associated with a constant increase in the molar oxygen/fuel ratio of 0.23. Increased AFR also diluted this enhanced mixing
with nitrogen, which may explain the decrease in slope with increased AFR.

2634 © 2019 by the Association for Iron & Steel Technology.


2.40E-07

Maximum CO Concentration
2.30E-07
2.20E-07
2.10E-07
(kmol/ft3) 2.00E-07
1.90E-07
1.80E-07
1.70E-07
1.60E-07
1.50E-07
16 17 18 19 20 21 22
AFR
Figure 4. CO concentration variation near the preheat top zone burner
Flame shape and temperature was also investigated. Figure 5 shows the side profile of the preheat zone top flames for each
AFR case. The temperatures found at the track point follow the same trend as the slab surface and bulk temperatures, wherein
the lowest temperature was found in the substoichiometric case and the highest in the slightly superstoichiometric case of
AFR 18. Of note was the high temperature zone in excess of 3400°F found in the 22 AFR case; this flame was extremely
premature and could damage the burner and surrounding refractory. This premature combustion was likely caused due to the
large air/fuel ratio. The track point temperature is still higher than the temperature found in the substoichiometric case. The
temperature for case 2 is notably higher than that of case 3 with a 128°F difference; this is most likely due to the longer flame
shape found in case 2.

Figure 5. Flame profiles in the top preheat zone with track point shown

Oxygen Enrichment
Oxygen-enriched combustion (OEC) has various noted benefits [11, 31-32]. With OEC, the total amount of nitrogen in the
furnace is reduced; this has the potential to lower nitrogen oxide (NOx) emissions. Other emissions may also be reduced, such
as carbon monoxide and hydrocarbons. OEC is also purported to increase efficiency as the mass of gases leaving the furnace
is reduced and less energy is required to bring nitrogen up to the temperature of the combustion products; radiative heat
transfer is also enhanced. Productivity can also see an increase when fuel flow rate is held constant due to higher flame
temperatures and increased heat transfer to the slabs. The stability of the combustion itself has also been found to improve
due to better mixing of the oxygen and the natural gas.
Liu et al. conducted a small parametric study to see the effects of oxygen enrichment on the furnace gas temperature [24]. The
volume fraction of oxygen within the combustion air was increased and decreased by 2% for a total of 3 cases. They found

© 2019 by the Association for Iron & Steel Technology. 2635


that the resulting gas temperature 3 inches above the slabs and in the centerline of the furnace increased about 2% with 2%
more oxygen but decreases around 4% with 2% less oxygen. Temperatures in certain places vary by as much as 100°F, which
has shown the possibility of enhanced heat transfer. Their results are shown in Figure 6, which shows the temperature trend
along the line 3 inches above the slabs along the centerline of the furnace.

2600

2500

2400
Gas Temperature (°F)

2300

2200

2100

2000
18.9% oxygen
1900 20.9% oxygen
22.9% oxygen
1800
0 10 20 30 40 50 60 70 80 90 100 110
Distance from Charge Door (ft)

Figure 6. Gas temperature at the center face of the furnace with different oxygen volume fraction (Liu et al.)
In the present study, a more in-depth and properly implemented parametric study has been conducted according to Table 4 in
alignment with the presented AFR results above; here, the ratio of total oxygen to fuel was held constant while the level of
oxygen enrichment of the air changed. Specifically, this necessitated decreasing the flow rate of the enriched air to ensure
that the oxygen fuel ratio stayed constant at the same fuel rate. The ‘base’ case represents air-fuel combustion, and cases 1, 2
and, 3 are a low oxygen enrichment level, medium oxygen enrichment level, and oxy-fuel respectively. For the present study,
only case 1 is discussed in comparison with the base case. Higher oxygen enrichment levels will be discussed at a future
time.

Table 4. Oxygen enrichment study operating conditions

Air Mass Flow Rate (ft3/s) Avg.


Case Molar Oxygen Oxygen/Fuel
Preheat Zone Preheat Zone Top
Bottom Ratio
Base 0.21 339.7 417.4
1 0.29 246.0 302.2
2.285
2 0.46 155.1 190.5
3 1.00 71.3 87.6

Oxygen enrichment for the low oxygen enrichment case was applied to the preheat zone, so the investigation of slab bulk
temperature and any scale formation insights are based around the residence time of the slab from charging to the end of
preheating. Figure 6 shows the predicted slab heating profile of the low oxygen enrichment case to the base stoichiometric
case. The temperature profiles are similar with the base line case staying slightly hotter than the oxygen enriched case for the
first hour of residence time. Near the end of the zone, the slab bulk temperature of the low oxygen enrichment case (case 1)
begins to overtake the base case.

2636 © 2019 by the Association for Iron & Steel Technology.


1600

1400

Slab bulk temperature (°F)


1200

1000

800

600

400

200 Low Oxygen Enrichment


Base
0
0 10 20 30 40 50 60 70
Residence time (min)
Figure 6. Average slab bulk temperature compared between the base case and the low oxygen enrichment case
Schluckner et al. found that the temperature of their samples as heated in oxygen enriched combustion atmospheres quickly
surpassed those in air-fuel combustion atmospheres [19]. The samples used by Schluckner were much smaller than those found
in the present study. With this is mind, the long time period taken for the slab bulk temperature of the low oxygen enrichment
case to exceed that of the air-fuel combustion case is to be expected. The furnace used in the current study also had a larger
domain and differing heat transfer effects compared to the stationary setup Schluckner et al. utilized. For the current study,
the total heat transfer into the slabs was reduced with oxygen enrichment; however, the percentage of heat transfer into the
slabs from radiation increased from 93.85% to 97.53%. This is somewhat significant as convection usually has a larger effect
within the preheat zone than the following zones due to the backflow from throughout the furnace to the outlets and the lower
slab temperatures found in this zone. The final bulk slab temperature at the end of the preheat zone for the low-level oxygen
enriched case was 30°F higher compared to the base case.
Assuming significant scale formation began when the slab surface exceeded a temperature of 1292°F (700°C), the scale
thickness is shown in Figure 7 as a function of time spent oxidizing while in the preheat zone. The average surface
temperature of the base case exceeded 1292°F first; as such, the length of time oxidizing was longer. After the first few
minutes, the slope of the oxidation between the 2 cases begins to diverge with the base case showing an overall larger rate of
oxidation.
While this may seem fortuitous in terms of less oxidized product with increased enrichment, it would seem anomalous in
comparison with other scale formation studies done through CFD analysis. Schluckner et al., for example, found that
oxidation increased with oxygen enrichment for a given residence time as attributed to the larger concentrations of oxidizing
species within the furnace atmosphere [19]. However, operational experience at industrial facilities where oxygen enrichment
has been implemented have not resulted in yield decreases due to oxygen enrichment [33]. As there are many variables within
the scale formation model used in this study, further analysis is required. At first blush, the increased scale formation in the
base case is assumed to be caused by the increased gas flow rate within the furnace and the resulting influence on the mass
transfer coefficient within the scale formation model. Both scale thicknesses are fairly close to one another and show a
difference of around 0.0046 inches, which is only 9.12% of the final predicted base case thickness (0.0504 in).

© 2019 by the Association for Iron & Steel Technology. 2637


1.20E-02

1.00E-02

Scale thickness (in)


8.00E-03

6.00E-03

4.00E-03

2.00E-03 Base
Low Oxygen Enrichment
0.00E+00
0 10 20 30 40 50
Time (min)
Figure 7. Scale thickness between a low oxygen enriched combustion case and an air-fuel combustion case (base)

CONCLUSIONS
A 3D CFD model of an industrial reheating furnace has been developed and used to investigate the effect of varied air/fuel
ratio and oxygen enrichment on combustion performance. Furthermore, a previously developed scale formation model has
been applied to investigate the effects of changing the composition and amount of combustion air going into the furnace on
scale formation [21]. As expected, varying air/fuel ratio (AFR) caused changes in predicted slab bulk temperature and in scale
formation. Up to a limit, an increase in air rate into the furnace at higher AFR showed an increase in extent of combustion
and predicted an increase in scale thickness. Slightly superstoichiometric operation is where the highest slab temperature was
found. This was attributed to improved mixing that led to more complete combustion. Above an estimated AFR of 20, the
benefits of mixing are outweighed by the excess nitrogen in the ballast air and cause the furnace gas and the subsequent slab
temperature to decrease.
An examination of oxygen enrichment was modelled to further investigate the effects of combustion on reheat furnace
productivity and scale formation. Results show that slight oxygen enrichment increased the slab bulk temperature by 30°F
compared to the base case. As the amount of fuel flowing into the furnace was the same for each case, a slight improvement
in combustion efficiency (in terms of the ratio of heat absorbed into the slab over the total furnace heat input) was noted.
Comparing scale formation as a function of residence time that the average slab surface temperature was above 1292°F
(700°C), the model predicted a decrease in the rate of scale formation in the case of oxygen enrichment compared to the base
case. The base case had larger scale growth due to the longer residence time above the oxidation temperature limit and the
higher gas velocity near the slab. Further oxygen enrichment studies are planned in order to investigate the impact of higher
levels of oxygen enrichment.

ACKNOWLEDGEMENTS
We would like to thank the Steel Manufacturing Simulation & Visualization Consortium (SMSVC) members for funding this
project as well as their technical feedback. We would also like to thank the members of the Center for Innovation through
Visualization and Simulation (CIVS) for their knowledge and input throughout the project.

REFERENCES
1. “Mineral commodity summaries 2018,” U.S. Geological Survey, 2018, p. 82.
2. S.H. Han, S.W. Baek, M.Y. Kim, “Transient radiative heating characteristics of slab in a walking beam type reheating
furnace,” International Journal of Heat and Mass Transfer, Vol. 52, No. 3-4, January 2009, pp. 1005-1011.
3. P.V. Barr, “The development, verification, and application of a steady-state thermal model for the pusher-type reheat
furnace,” Metallurgical and Materials Transactions B, Vol. 26, No. 4, August 1995, pp. 1543-1916.

2638 © 2019 by the Association for Iron & Steel Technology.


4. L. Tang, H. Ren, Y. Yang, “Reheat furnace scheduling with energy consideration,” International Journal of Production
Research, Vol. 53, No. 6, June 2014, pp.1642-1660.
5. S.H. Han, D. Chang, C. Huh. “Efficiency analysis of radiative slab heating in a walking-beam-type reheating furnace,”
Energy, Vol. 36, February 2011, pp. 1265-1272.
6. M. Si, S. Thompson, K. Calder. “Energy efficiency assessment by process heating assessment and survey tool (PHAST)
and feasibility analysis of waste heat recovery in the reheat furnace at a steel company,” Renewable and Sustainable
Energy Reviews, Vol. 15, No. 6, August 2011, pp. 2904-2908.
7. H.E. Pike, S.J. Citron, “Optimization studies of a slab reheating furnace,” Automatica, Vol. 6, No. 1, January 1970, pp.
41-50.
8. W.H. Chen, Y.C. Chung, J.L. Liu, “Analysis on energy consumption and performance of reheating furnaces in a hot
strip mill,” International Communications in Heat and Mass Transfer, Vol. 32, No. 5, April 2005, pp. 695-706.
9. A. Steinboeck, D. Wild, T. Kiefer, A. Kugi, “A mathematical model of a slab reheating furnace with radiative heat
transfer and non-participating gaseous media,” International Journal of Heat and Mass Transfer, Vol. 53, No. 25-26,
December 2010, pp. 5933-5946.
10. F.A.D. Oliveria, J.A. Carvalho Jr., P.M. Sobrinho, A. de Castro. “Analysis of oxy-fuel combustion as an alternative to
combustion with air in metal reheating furnaces,” Energy, Vol. 78, December 2014, pp. 290-297.
11. C.E. Baukal Jr., “Basic Principles,” Oxygen-Enhanced Combustion, Taylor & Francis, Boca Raton, FL, 1998, pp. 8-51.
12. Z. Li, P. V. Barr, J. K. Brimacombe. “Computer Simulation of the Slab Reheating Furnace,” Chemical and Extractive
Metallurgy, Vol. 27, No. 3, 1988, pp. 187-196.
13. H.J. Karimi, M.H. Saidi, “Heat transfer and energy analysis of a pusher-type reheating furnace using oxygen enhanced
air for combustion,” Journal of Iron and Steel Research International, Vol. 17, No. 4, April 2010, pp. 12-17.
14. S.H. Han, Y.S. Lee, J.R. Cho, K.H. Lee. “Efficiency analysis of air-fuel and oxy-fuel combustion in a reheating
furnace,” International Journal of Heat and Mass Transfer, Vol. 121, June 2018, pp. 1364-1370.
15. M.Y. Kim, “A heat transfer model for the analysis of transient heating of the slab in a direct-fired walking beam type
reheating furnace,” International Journal of Heat and Mass Transfer, Vol. 50, No. 19-20, September 2007, pp. 3740-
3748.
16. R. Prieler, B. Mayr, M. Demuth, B. Holleis, C. Hochenauer, “Prediction of the heating characterisics of billets in a
walking hearth type reheating furnace using CFD,” International Journal of Heat and Mass Transfer, Vol. 92, January
2016, pp. 675-688.
17. R. Prieler, B. Mayr, M. Demuth, B. Holleis, C. Hochenauer, “Numerical analysis of the transient heating of steel billets
and the combustion process under air-fired and oxygen enriched conditions,” Applied Thermal Engineering, Vol. 103,
June 2016, pp. 252-263.
18. B. Mayr, R. Prieler, M. Demuth, L. Moderer. “CFD modelling and performance increase of a pusher type reheating
furnace using oxy-fuel burners,” Energy Procedia, Vol. 120, August 2017, pp. 462-469.
19. C. Schluckner, C. Gaber, M. Demuth, S. Forstinger, R. Prieler, C. Hochenauer. “CFD-model to predict the local and
time-dependent scale formation of steels in air- and oxygen enriched combustion atmospheres,” Applied Thermal
Engineering, Vol. 143, October 2018, pp. 822-835.
20. J.H. Jang, D.E. Lee, M.Y. Kim, H.G. Kim. “Investigation of the slab heating characteristics in a reheating furnace with
the formation and growth of scale on the slab surface,” International Journal of Heat and Mass Transfer, Vol. 53, No.
19-20, September 2010, pp. 4326-4332.
21. X. Liu, B. Worl, G. Tang, A.K. Silaen, J. Cox, K. Johnson, C.Q. Zhou. “A Numerical Model to Predict Scale Formation
in an Industrial Reheat Furnace,” Steel Research International, October 2018, pp. 62-70.
22. L. Himmel, R.F. Mehl, C.E. Birchenall. "Self-diffusion of iron in iron oxides and the Wagner theory of
oxidation." JOM, Vol. 5, No. 6, June 1953, pp. 827-843.
23. H.F. Marston, P.H. Bolt, G. Leprince, M. Röder, R. Klima, J. Niska, M. Jarl. “Challenges in the modelling of scale
formation and decarburisation of high carbon, special and general steels,” Ironmaking and Steelmaking, Vol. 31, No. 1,
2004, pp. 57-65.

© 2019 by the Association for Iron & Steel Technology. 2639


24. X. Liu, G. Tang, A. Silaen, J. Cox, K. Johnson, C. Zhou. “Investigation of Operational Effects on Slab Heating
Characteristics in a Pusher-Type Reheat Furnace with Continuous Slab Motion,” ASME 2017 Heat Transfer Summer
Conference, Vol. 1, July 2017.
25. J.G. Kim, K.Y. Huh, I.T. Kim, “Three-dimensional Analysis of the Walking-beam-type Slab Reheating Furnace in Hot
Strip Mills,” Numerical Heat Transfer, Part A: Applications, Vol. 38, No. 6, 2000, pp. 589-609.
26. C.K. Westbrook, F.L. Dryer. “Simplified Reaction Mechanisms for the Oxidation of Hydrocarbon Fuels in Flames,”
Combustion Science and Technology, Vol. 27, No. 1-2, 1981, pp. 31-43.
27. C. Yin, L.A. Rosendahl, and S.K. Kær. “Chemistry and radiation in oxy-fuel combustion: A computational fluid
dynamics modeling study,” Fuel, vol. 90, no. 7, July 2011, pp. 2519-2529.
28. R.Y. Chen, W.Y.D. Yuen. “Review of the High-Temperature Oxidation of Iron and Carbon Steels in Air or Oxygen,”
Oxidation of Metals, Vol. 59, No. 5-6, June 2003, pp. 433-468.
29. J. Akiyama, H. Ohta, R. Takahashi, T. Waseda, J. Yagi. “Measurement and Modeling of Thermal Conductivity for
Dense Iron Oxide and Porous Iron Ore Agglomerates in Stepwise Reduction,” ISIJ International, Vol. 32, No. 7,
January 1992, pp. 829-837.
30. B. Jenkins, P. Mullinger. “Industrial and Process Furnaces: Principles, Design and Operation,” Elsevier Ltd.,
Burlington, MA, 2008, pp. 61-317.
31. DOE and Industrial Heating Equipment Association. “Improving Process Heating System Performance: A Sourcebook
for Industry,” DOE and Industrial Heating Equipment Association (IHEA), 2005. [Online]. Available:
www.oit.doe.gov/bestpractives/library.shtml.
32. Energy Center of Wisconsin. “Oxygen-Enriched Combustion Technologies,” Energy Center of Wisconsion, Publication
no. 426-1, 2000. [Online].
33. K. Tian. Personal communication, Praxair, 2018-2019.

2640 © 2019 by the Association for Iron & Steel Technology.

You might also like