You are on page 1of 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/263590998

MULTIPHASE FLOW AND LOW PRESSURE EFFECTS IN THE SEN

Conference Paper · June 2014

CITATIONS READS

0 420

2 authors:

Mirko Javurek Maria Thumfart


Johannes Kepler University Linz Johannes Kepler University Linz
50 PUBLICATIONS   246 CITATIONS    6 PUBLICATIONS   21 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

K1-Met: Magnetohydrodynamics in Metallurgical Processes View project

All content following this page was uploaded by Mirko Javurek on 03 July 2014.

The user has requested enhancement of the downloaded file.


MIRKO JAVUREK1, MARIA THUMFART1

MULTIPHASE FLOW AND LOW PRESSURE EFFECTS IN THE SEN

Abstract

In the continuous casting of steel a stopper rod is commonly used to control the flow
rate from the tundish into the mould. Agglomeration of solid material near the stopper rod and
inside the submerged entry nozzle (SEN) can lead to a reduced cross section and thus to a
decreased casting speed or even total blockage (“clogging”). It is known that the injection of
argon gas at the stopper rod tip reduces the clogging effect. Nevertheless the mechanisms
involved in clogging are still not fully understood. A key for a better understanding may be the
consideration of the absolute pressure: Single phase considerations of the flow in the region of
the stopper rod result in a low or even negative pressure at the smallest cross section due to
the high flow velocities. This can cause degassing of dissolved gases from the melt,
evaporation of alloys and entrainment of air through the refractory material. According to
these considerations, a two phase liquid-gas flow is expected even in the case when no gas is
injected in the stopper-rod region. A mathematical model for this two phase flow situation in
the SEN is presented that links the throughput and the pressure inside of the SEN with the
geometrical parameters, the amount of gas in the SEN and the pressure loss coefficient at the
stopper rod. An explanation theory for the diminishing influence of the gas injection on
clogging is outlined.

Keywords
Submerged entry nozzle (SEN), clogging, partially filled SEN, stopper rod, argon injection,
cavitational effects

1. Introduction

In the continuous casting process a stopper rod is a commonly used device to control the
steel flow from the tundish to the mould. During the casting process solid material mainly
consisting of oxides tends to agglomerate at the stopper and the submerged entry nozzle
(SEN) generally known as clogging [1-3]. This can lead to a reduced throughput or even to a
total blockage of the SEN as well as severe holdup in the casting process.
Clogging is the end product of a wide range of complex mechanisms such as chemical
reactions, heat transfer, particle transport and deposition, degassing, diffusion, bubble
formation and bubble transport [2], which influence each other. The chemical reactions can
produce gas bubbles and inclusions. Gas bubbles in the flow can trigger chemical reactions,
collect inclusions and change large scale flow characteristics. Inclusions can nucleate bubbles
or agglomerate at the walls and thus change their properties. The amount and distribution of
gas in the liquid steel highly influences the flow field and thus all other phenomena included
in clogging as well. So an important step towards understanding the flow situation near the
stopper rod is to understand the two phase flow. Nevertheless we start with a consideration of
the single phase flow situation before we take a closer look at the two phase flow.

1
Johannes Kepler University Linz, Institut für Strömungslehre und Wärmeübertragung
2. Single phase state model

Fig. 1 shows the pressure along a streamline from the tundish surface to the outflow
ports of the SEN considering a single phase flow. The pressure is plotted as a function of the
vertical coordinate for liquid steel and water (1:1 scale water model). Bernoulli‘s equation can
be used to describe the flow from the surface of the tundish to the narrowest point in the
stopper rod gap as no flow separation occurs. The pressure at the surface of the tundish is
equal to the atmospheric pressure. Then it rises due to the increasing hydrostatic pressure.
Near the bottom of the tundish the flow acceleration of the fluid causes a pressure drop. It
reaches a minimum at the point of the smallest cross section. Underneath the stopper a free jet
forms [4], which causes the main pressure loss. In the SEN the pressure rises again due to
gravity.
Assuming a single phase flow the absolute pressure in the stopper rod gap reaches
negative values for most typical geometrical configurations. Negative pressure contradicts the
common concept of pressure as a representation of collision force of the molecules.
Nevertheless a negative pressure can occur in very clean liquids. Theoretical considerations
predict a tensile strength of water as low as −10 000 bar due to the Van-der-Waals forces [5].
Less than −200 bar were already measured in water experiments [5]. This tensile strength can
be observed in pure liquids where the near order of the molecules is not disturbed. Any

Fig. 1: single phase SEN flow, pressure along a Fig. 2: two phase SEN flow, pressure along a streamline
streamline from the tundish surface to the outflow ports from the tundish surface to the outflow ports of the SEN
of the SEN for steel (black line) and water (grey line) for steel (black line) and water (grey line)
discontinuity in the near order can be the starting point of rupture and thus gas accumulation.
This process is called nucleation. Depending on the surface tension of the liquid, a certain
critical size of the nuclei has to be reached in order to enable gas bubbles to grow. Due to the
high surface tension of liquid steel, these nuclei need to be much larger than in water [6,7].
While cavitation in water can take place in the whole liquid due to the dense distribution of
micro-particles (order of magnitude 1 µm), cavitational effects in liquid steel can only arise at
rough SEN and stopper rod wall pores or at very dilute large non-metallic particles (all of
them order of magnitude 100 µm) [6,7]. Without the extra addition of inert gas, the formation
of gas curtains at the SEN and stopper rod walls filled with gasses dissolved from the liquid
steel (Nitrogen, Oxygen and Hydrogen) and possibly sucked gasses from the surrounding
atmosphere is expected [6,7].
Besides the formation of gasses due to low pressures, inert gas is often injected near the
stopper rod gap in order to reduce clogging. Thus a general two-phase flow situation is
considered in the following with liquid steel and a gas phase, covering both types of gas
(injected and nucleated).

3. Two-Phase State Model

The above presented single phase state model is now extended to a two phase model by
adding a gas phase to the primary liquid phase. The phase distribution inside the SEN is
considered in the most general way possible. Therefore, it is assumed that the gas can not only
form bubbles inside the liquid steel, but also other regimes with a high gas volume fraction
like gas cavities with steel droplets inside as sketched in fig. 3. It is first assumed that the
lower part of the SEN is filled by liquid steel up to the height h2 over the mould level (see fig.
3). Based on this assumption the pressure distribution can be determined as sketched in fig. 2.

Fig. 3: pressure along a streamline from the tundish surface to the outflow ports of the SEN for
steel (black line) and water (grey line)

Starting at the mould level the pressure inside the SEN drops with increasing height due
to the hydrostatic pressure until the phase interface steel/gas is reached at h2. From there on,
the pressure stays constant at level p2. Bernoulli’s equation for a streamline between the phase
interface and the mould level is
p2  gh2  12 u22  p0   W 12 u22   P 12 u22 (1)
5
where p0 is the ambient atmospheric pressure (about 10 Pa), ρ is the liquid steel density, g is
the gravitational acceleration, u2 is the mean flow velocity in the SEN,  W is the pressure loss
coefficient due to the SEN wall friction and  P is the pressure loss coefficient of the
submerged free jets leaving the entry nozzle through the nozzle ports. From this equation
p2  p0  gh2   W   P  1 12 u22 . (2)
It is expected that there is a 100 % pressure loss (  P  1 ) and thus the dynamic pressure term
at the right side of equation 2 is reduced to the comparably low wall friction  W .
In the case that there is a more smeared phase distribution (e.g. due to gas bubbles in the
liquid phase, liquid droplets in the gas phase and due to foam-like states), h2 can be seen as an
equivalent SEN filling height (the equivalent filling height for a situation with sharply
separated phases produces the same pressure p2 as in the case of a smeared interface).
Starting from the tundish level the pressure inside the tundish increases with decreasing
height due to the hydrostatic pressure until the tundish bottom wall is nearly reached. Here the
pressure is
p1  p0  gh1
(3)
where h1 is the tundish fill height. As the liquid steel leaves the tundish, it starts to accelerate
and flows into the gap between stopper rod and SEN.
The pressure drops until pmin when the highest flow velocity u is reached in the
narrowest cross section of the gap. Thus
pmin  p1  12 u 2 (4)
where the rightmost term is the dynamic pressure drop calculated with the gap flow velocity u.
The gap flow velocity is related to the SEN flow velocity u2 due to the conservation of mass
for steady conditions by
uAS  u2 A2
(5)
where AS is the total gap cross section area and A2 is the SEN cross section area.
After the narrowest cross section of the gap, the cross section expands, the flow
decelerates and regains pressure until the flow separates from the SEN and stopper rod walls,
reaching the pressure p2. Bernoulli’s equation for a streamline between the tundish bottom and
the SEN top region is
p1  p2  12 u22   S 12 u 2
(6)
where  S is the pressure loss coefficient of the gap flow. If  S = 100 % no pressure is
regained and p2 = pmin. For the hypothetical case that  S is zero, the complete pressure would
be regained and p2 ≈ p1. The observed pressure loss coefficients from water model trials at
voestalpine are between 50% and 70%. The pressure loss coefficient depends mainly on the
gap geometry and only slightly on the Reynolds number and the stopper rod lift.
Inserting equation 2, 3 and 5 in equation 6 gives the gap flow velocity u for a given
SEN and tundish fill level
2 p1  p2  2 g h1  h2 
u 
  (7)
where the total pressure loss coefficient
AS2
   S   P   W 
A22 . (8)
Usually, the total pressure loss coefficient is dominated by the pressure loss coefficient  S of
the gap flow: for a typical situation as described below
2
 P   W  AS2  9 %
A2 . (9)
For the calculation of the overall throughput
2 g h1  h2 
m  uAS (hS )  C ' hS
 , (10)
the total gap cross section area AS has to be known as a function of the stopper rod lift hS. For
usual geometries AS depends linearly on the stopper rod lift hS with the proportionality factor
C’ [6].
Equation 4 in combination with equation 6 and 10 gives the minimal pressure in the gap
 h h 
pmin  p0  g  h1  1 2 
  . (11)
Now the derived equations for the throughput and minimal pressure depend basically on
the following parameters: tundish level h1, SEN fill level h2, total pressure loss coefficient ζ
and stopper rod lift hS. It is remarkable that the minimal pressure in the gap does not directly
depend on the stopper rod lift. Nevertheless the total pressure loss coefficient slightly depends
on the stopper rod lift due to the SEN wall friction (equation 8) and probably also due to the
change of the gap geometry.
In fig. 4 the pressure inside the SEN p2 near the stopper rod tip (dashed line,
approximately independent of the pressure loss coefficient, equation 2) and the minimum
pressure in the gap pmin (solid black lines, equation 11) are plotted as a function of the fill
level h2 for different pressure loss coefficients ζ and constant stopper rod lift and tundish
level. Typical values and ranges for a setup used at the voestalpine slab casting steel plant in
Linz with a tundish level h1 = 1 m are applied. The grey lines in fig. 4 denote the
corresponding steel mass flow rates also for different pressure loss coefficients according to
equation 10 with C ' hS  7.07 kg m -1 .
In the case of an SEN completely filled with liquid steel (h2 = 0.84 m), the absolute
pressure p2 is 0.4 bar near the stopper rod tip and increases if the SEN is partially filled with
gas until it reaches 1.1 bar if the SEN level is located at the SEN ports. In this extreme case an
overpressure inside the SEN is necessary to push the SEN fill level below the mould level.
The minimal pressure in the gap pmin is lower or equal to the SEN pressure p2, depending on
the pressure loss coefficient: the lower the pressure loss coefficient, the lower the gap
pressure. For high SEN fill levels and lower pressure loss coefficients, low or even negative
gap pressure values pmin can be reached. For an increasing gas volume inside the SEN and for
an increasing pressure loss the throughput becomes lower.
In fig. 4 the corresponding data from a typical operating point of a full scale slab caster
water model of voestalpine is marked by circles (WM): the SEN is completely filled with
liquid and the water throughput for a stopper rod lift corresponding to the data plotted in fig. 4
corresponds to 3.1 t/min steel. According to equation 10 this gives a total pressure loss
coefficient of 68 % and – in the case of liquid steel – would result in a minimal pressure of
−0.18 bar.
From plant measurements it is supposed that the stopper rod lift is higher for the same
volumetric throughput than in the water model. This cannot be explained if the flow situation
was the same in the water model and the plant since the density cancels out and the kinematic
viscosity is almost the same for water and liquid steel. Furthermore a minimal pressure of
−0.15 bar may cause some degassing effects as mentioned above. The higher stopper rod lift
at the plant can be explained and the negative pressure can be avoided if the impact of gases is
considered as described in the following.

Fig. 4: pressure along a streamline from the tundish surface to the outflow ports of the
SEN for steel (black line) and water (grey line)

Our hypothesis is that the presence of gases – be it from degassing effects of from the
injection of gas – has two effects:
1. the pressure loss coefficient is increased and
2. the SEN is partially filled with gas.
Both effects reduce the throughput and increase the minimum pressure in the gap. The arrows
in fig. 4 with the dashed lines indicate how different combinations of the two effects can
increase the minimum pressure towards positive values and at the same time decrease the
mass flow. If degassing effects at low steel pressures were responsible for the clogging of the
SEN due to the formation of oxide particles, this could also explain why the injection of argon
gas reduces clogging.
The increase of the pressure loss coefficient by the presence of injected gas is expected
to arise because of gas that is transported by the flow into the gap between SEN wall and
stopper rod, which can also be observed at the water model (fig. 5). As a consequence the
flow separation is nearer to the smallest cross section between stopper rod and SEN, resulting
in a higher pressure loss. The effect is expected to be much higher in the case of the real plant
since the absolute pressure in the gap is much lower than in the water model and thus the
specific gas volume and the displacement is much higher. It is expected that the multiphase
flow and the phase distribution are highly unsteady (see also [8]).

Fig. 5: full scale water model at voestalpine with transparent SEN walls and stopper rod: bubbles from a gas
injection at the stopper rod tip rod sucked into the gap between stopper rod and SEN wall; black lines denote the
contours of wall and stopper rod

4. Conclusions

The model of a single phase flow of a completely filled SEN results in very low or even
negative pressure in the vicinity of the gap between stopper rod and SEN walls. While pure
liquids are able to stand certain negative pressures, in “real” liquids like water and liquid steel
degassing starts by the formation of gas bubbles at impurities (nuclei). Besides the formation
of gasses due to low pressures, inert gas is often injected near the stopper rod gap in order to
reduce clogging. Thus a general two-phase flow situation is considered with liquid steel and a
gas phase, covering both types of gas. The effect of different SEN gas filling levels and
pressure losses at the stopper rod gap are analysed: for high SEN fill levels and lower pressure
loss coefficients, low or even negative gap pressure values can be reached. For an increasing
gas volume inside the SEN and for an increasing pressure loss the minimal gap pressure
increases. The presence of a gas phase (originating from gas injection or degassing) is
supposed to have two effects: the pressure loss coefficient is increased due to an earlier flow
separation and/or the SEN is partially filled with gas. Both effects diminish the low pressure
situation. If degassing effects at low steel pressures were responsible for the clogging of the
SEN due to the formation of oxide particles, this could also explain why the injection of argon
gas reduces clogging.
By means of the presented results, the flow situation inside of the SEN and in the
vicinity of the stopper rod tip can be better understood. The developed mathematical model is
able to illustrate how throughput and pressure depend on the SEN gas fill level. In
combination with experimental results from water models and experiences at the plant, a
partial filling of the SEN with gas seems obvious for a wide range of operating conditions.
Intentionally the model is kept simple in order to be as comprehensive as possible. It can
easily be enhanced by considering currently neglected effects like the change of the pressure
loss coefficient as a function of the stopper rod lift.

References

[1] Kojola, N.; Ekerot, S.; Andersson, M.; Jönsson, P. G.; Pilot plant study of nozzle clog-
ging mechanisms during casting of REM treated stainless steels; Ironmaking and
Steelmaking, Vol. 38, Iss 1 (2011), P. 1-11

[2] Rachers K.G.; Thomas B. G.: Clogging in continuous casting nozzles.; In 78th Steel-
making Conference Proceedings; volume 78 (1995); P 732–734

[3] Zhang, L.; Wang, Y.; Zuo, X.; Flow Transport and Inclusion Motion in Steel Continu-
ous-Casting Mold under Subermged Entry Nozzle Clogging Condition; Metallurgical
and Materials Transactions B, Vol. 39B, Aug. 2008, P. 534-550

[4] Javurek, M.; Thumfart, M.; Wincor, R.: Investigations on Flow Pattern and Pressure in-
side SEN below Stopper Rod; steel research international, Vol. 81, Iss. 8 (2010), P. 668-
674

[5] Brennen, C. E.: Cavitation and Bubble Dynamics; Oxford University Press; Oxford;
(1995)

[6] Thumfart, M.; Javurek M.: Low pressure effects in SEN-stopper region in continuous
casting. Steel research international, accepted paper to be published 2014

[7] Thumfart, M.: Liquid steel at low pressure: Experimental investigation of a water air
flow in a convergent divergent nozzle. Proceedings ECCC 2014

[8] Zhe WANG. Kusuhiro MUKA, Daisuke IZU: Influence of Wettability on Fluid Flow
inside the Nozzle the and Behavior of Argon Bubbles and Mold. ISIJ International. Vol.
39 (1 999), No. 2, P. 154-163

View publication stats

You might also like