Processing math: 0%

Effect of gas blowing nozzle angle on multiphase flow and mass transfer during RH refining process

Jiahao Wang, Peiyuan Ni, Chao Chen, Mikael Ersson, Ying Li

图(16)  /  表(1)

分享

计量
  • 文章访问数:  1016
  • HTML全文浏览量:  292
  • PDF下载量:  65
  • 被引次数: 16

目录

Cite this article as:

Jiahao Wang, Peiyuan Ni, Chao Chen, Mikael Ersson, and Ying Li, Effect of gas blowing nozzle angle on multiphase flow and mass transfer during RH refining process, Int. J. Miner. Metall. Mater., 30(2023), No. 5, pp.844-856. https://dx.doi.org/10.1007/s12613-022-2558-5
Jiahao Wang, Peiyuan Ni, Chao Chen, Mikael Ersson, and Ying Li, Effect of gas blowing nozzle angle on multiphase flow and mass transfer during RH refining process, Int. J. Miner. Metall. Mater., 30(2023), No. 5, pp.844-856. https://dx.doi.org/10.1007/s12613-022-2558-5
引用本文 PDF XML SpringerLink
研究论文

RH精炼过程气体喷吹角度对多相流动及传质的影响

文章亮点

(1) 提出并研究了旋转流驱动RH高效精炼新技术 (2) 系统揭示了旋转喷吹对提升循环流量和缩短混匀时间的作用机理 (3) 全面掌握了喷吹角度对RH高效精炼过程多相流场的影响规律
当前,在我国制造业高质量发展和“双碳”战略下,钢铁工业的高端、绿色、低碳、高效率的创新发展十分关键。RH精炼具有脱碳、脱气、脱硫、去除夹杂物及调整钢水成分等冶金功能,已成为轴承钢等诸多高品质钢生产的重要工序之一。本研究工作基于提高RH上升管气泡分散度、扩大气泡与钢水相互作用范围的思路,创新性的提出了旋转喷吹新技术,通过简单改变上升管气体喷吹角度,实现上升管内钢水的旋转流动,而钢水旋转流动进一步促进气泡的弥散分布,从而充分发挥气泡对钢水的提升作用,提高上升管横截面速度分布的均匀性和RH精炼的循环流量。本文首先开展了传统喷吹条件下的数值模拟和水模型实验研究,结果表明,气体流量为40 L·min-1时,数值模型预测的混匀时间与水模型实验测得的结果吻合较好,在选定的三个监测点位置,模型预测值与水模型实测值的误差在1.3%~7.3%范围内,证明了数值模型的可靠性;其次,开展了不同气体喷吹角度对RH精炼过程多相流动及传质行为的影响,研究表明,传统喷吹角度下的循环流量为9 kg·s-1,当喷嘴水平旋转30°或45°时,循环流量提高了约15%,此外,预测的三个监测点的混匀时间分别缩短了约21.3%、28.2%和12.3%。喷吹角度为30°和45°时,对128个气泡的统计分析表明,气泡在流体中的平均停留时间增加了约33.3%。研究表明,新的喷吹技术有利于提升RH循环流量和混匀时间,有望为进一步提升RH精炼效率提供了重要解决思路。

 

Research Article

Effect of gas blowing nozzle angle on multiphase flow and mass transfer during RH refining process

Author Affilications
    Corresponding author:

    Peiyuan Ni      E-mail: nipeiyuan@smm.neu.edu.cn

    Ying Li      E-mail: liying@mail.neu.edu.cn

  • Received: 18 August 2022; Revised: 15 September 2022; Accepted: 28 September 2022; Available online: 29 September 2022
A three-dimensional mathematical model was developed to investigate the effect of gas blowing nozzle angles on multiphase flow, circulation flow rate, and mixing time during Ruhrstahl-Heraeus (RH) refining process. Also, a water model with a geometric scale of 1:4 from an industrial RH furnace of 260 t was built up, and measurements were carried out to validate the mathematical model. The results show that, with a conventional gas blowing nozzle and the total gas flow rate of 40 L·min–1, the mixing time predicted by the mathematical model agrees well with the measured values. The deviations between the model predictions and the measured values are in the range of about 1.3%–7.3% at the selected three monitoring locations, where the mixing time was defined as the required time when the dimensionless concentration is within 3% deviation from the bath averaged value. In addition, the circulation flow rate was 9 kg·s–1. When the gas blowing nozzle was horizontally rotated by either 30° or 45°, the circulation flow rate was found to be increased by about 15% compared to a conventional nozzle, due to the rotational flow formed in the up-snorkel. Furthermore, the mixing time at the monitoring point 1, 2, and 3 was shortened by around 21.3%, 28.2%, and 12.3%, respectively. With the nozzle angle of 30° and 45°, the averaged residence time of 128 bubbles in liquid was increased by around 33.3%.

 

  • Ruhrstahl-Heraeus, namely RH furnace, is one of the most important refining equipment in steel production, mainly used for some high-quality steel production such as ultra-low carbon steel, electrical steel, bearing steel, pipeline steel, and so on. The structure of a RH furnace is shown in Fig. 1. At present, RH furnace has been developed into multifunctional refining equipment with decarburization, degassing, desulfurization, the removal of inclusions, and the adjustment of steel temperature and compositions.

    Figure  1.  Schematic diagram of a RH furnace.

    Mixing time and circulation flow rate are the key characteristics to evaluate the refining efficiency of RH furnace, which are significantly influenced by steel flow behaviors. Due to the limitation of the high steel temperature, a direct investigation on multiphase flow behaviors during a RH refining process is very difficult. Therefore, numerical simulations have been widely used. Initially, the flow phenomena during a RH refining were simulated to be simplified into a single-phase flow [16]. Therefore, it cannot properly describe the multi-phase flow phenomena which happen in a real RH reactor. Then, the Eulerian-Lagrangian [7] and Eulerian-Eulerian [89] approach were used to simulate the RH refining process. These two methods can generally depict the flow phenomena, which is validated by water model experiments. In above studies, the vacuum chamber and the top surface of the ladle were assumed to be flat, ignoring the existence of top gas or top slag phase. In order to track the interface behavior, a combined discrete phase model (DPM)–volume of fluid (VOF) approach was used to simulate the RH refining process [1011]. Recently, a new model called the particle-free surface coupled model was used to investigate the argon-steel-slag flow in the vacuum chamber [12], where the size change of gas bubbles was considered. Chen et al. [13] and Shao et al. [14] established a coupled model of computational fluid dynamics model and population balance model to simulate the multiphase flow during RH refining process. Numerical simulations are usually validated either by water model experiments or by industrial measurements to show its reliability [6,10].

    The circulation flow rate and mixing time of a RH refining process depend on many variables. Gas blowing nozzle number is an important factor that affects the circulation flow rate. Generally, an increased number of gas plumes in the up-snorkel leads to an increased steel volume influenced by the injected gas, and an improved decarburization efficiency [1517]. However, the circulation flow rate can be decreased when too many nozzles are used [10]. This is due to that each nozzle plume coalesces into a large plume. Gas blowing flow rate is another important factor that affect the circulation flow rate, which has been extensively studied [10,15,1820]. It was found that the circulation flow rate can be improved by an increased gas flow rate until a ‘saturated level’ was reached. In addition, a lower vacuum chamber pressure can obtain a larger circulation flow rate [2124]. However, the increase in the gas blowing flow rate is supposed to be more efficient than a low vacuum degree [25]. The multi-leg design and an oval or arched shape of the snorkel cross section was found to be effective to enhance the circulation flow rate [2630]. Also, a ladle bottom gas injection can shorten the mixing time and improve the refining efficiency [3134]. However, an excessive gas flow rate can lead to the formation of a slag eye in the ladle and the accelerate erosion on the up-snorkel [3536].

    The circulation flow in a RH system was driven by the gas injection, resulting from the effect of gravity and buoyancy force. The velocities of both liquid and gas in the up-snorkel are mainly in the vertical direction. An additional horizontal velocity may help to improve the mixing of the liquid during passing by the up-snorkel. Li et al. [37] investigated the rotational flow phenomena generated by using a mechanical stirring impeller inside the up-snorkel. Water model experiments showed that this technique can increase the circulation flow rate by about 25%. In addition, electromagnetic stirring method was used to produce a rotational flow in the up-snorkel [3839]. The results showed that a large circulation flow rate can be obtained with the effect of swirling flow. Qi et al. [40] found that the generation of a rotational flow can increase the residence time of bubbles. In addition, Wang et al. [41] applied a forward-reverse rotational electromagnetic field to suppress the central vortex formation and to enhance the bubble refinement.

    In this work, a new method to produce a rotational flow in the up-snorkel was investigated and it could realize by simply changing the gas blowing nozzle direction. Both water model experiments and numerical simulations were carried out to investigate the multiphase flow behavior in the new design. The effect of the gas blowing nozzle angle on the multiphase flow, mixing time, and circulation flow rate were investigated.

    A water model with a geometric scale of 1:4 from a 260 t industrial RH furnace was built up. The dimensions of the water model are given in Table 1. Both the water model and the prototype have 16 nozzles, where 8 nozzles are uniformly distributed on the circumference of an up-snorkel cross section. The specific layout of the nozzles is shown in Fig. 2. The snorkels are immersed 110 mm beneath the free surface in ladle. The similarity of the two systems was guaranteed by the same value of Froude number in Eqs. (1) and (2). The gas flow rate in the water model can be obtained by using Eq. (3) [43], with the value of 40 L·min–1. The pressure in the vacuum chamber is around 97.7 kPa.

    Table  1.  Parameters of water model geometry and fluid physical properties
    Height
    of ladle /
    mm
    Diameter
    of ladle top /
    mm
    Diameter
    of ladle
    bottom / mm
    Liquid
    level of
    ladle / mm
    Diameter
    of snorkel /
    mm
    Length
    of snorkel /
    mm
    Distance
    between
    snorkel / mm
    1035978929800180412.5394.5
    Diameter
    of vacuum
    chamber / mm
    Number
    of
    nozzles
    Diameter
    of nozzles /
    mm
    Lower layer nozzle
    height from the bottom
    of snorkel / mm
    Distance between
    two rows of
    nozzles / mm
    Density of water /
    (kg·m–3)
    Viscosity
    of water /
    (10–3 Pa·s)
    574.5162505010001.003
    Density of
    air / (kg·m–3)
    Viscosity
    of air /
    (10–5 Pa·s)
    Density of the
    saturated NaCl solution /
    (kg·m–3) [42]
    Molar concentration
    of the saturated NaCl
    solution / (mol·L–1) [42]
    Interfacial tension
    of water and
    air / (N·m–1)
    1.2251.791197.25.3260.07197
    下载: 导出CSV 
    | 显示表格
    Figure  2.  Top view of nozzle arrangements (a) conventional nozzle and (b) rotated nozzle.

    The tracer addition position and the monitoring points in the water model are shown in Fig. 3. The stimulus-response approach was used to measure the mixing time in the bath. When the flow in the whole system is fully developed (about 270 s after the start of the gas blowing), 200 mL NaCl saturated solution was poured into the vacuum chamber as the tracer. Then, the changes of the electrical conductivity with time at different locations were measured to determine the bath mixing time.

    Figure  3.  Locations of monitor points, tracer addition, and different cross sections.
    F r = \frac{{{\rho _{\text{g}}}{{\boldsymbol{v}}_{\text{g}}}^2}}{{gL\left( {{\rho _{\text{l}}} - {\rho _{\text{g}}}} \right)}} (1)
    \lambda = \frac{{{L_{\text{m}}}}}{{{L_{\text{p}}}}} (2)
    \frac{{{Q_{\text{m}}}}}{{{Q_{\text{p}}}}} = {\lambda ^{2.5}}{\left( {\frac{{{\rho _{{\text{lm}}}}}}{{{\rho _{{\text{lp}}}}}} \times \frac{{{\rho _{{\text{gp,298 K}}}}}}{{{\rho _{{\text{gm}}}}}} \times \frac{{{P_{\text{m}}}}}{{{P_{\text{p}}}}} \times \frac{{{T_{\text{p}}}}}{{{T_{\text{m}}}}}} \right)^{0.5}} (3)

    where Fr is the Froude number; ρg is gas density, kg·m–3; ρl is liquid density, kg·m–3; g is the gravitational acceleration rate, m·s–2; L is a feature length, m; vg is gas velocity, m·s–1; λ is the geometric similarity ratio; Q is gas flow rate, m3·s–1; P is gas pressure, Pa; T is temperature, K; the subscript m and p represents the water model and prototype, respectively.

    The mathematical model was built up based on the following assumptions: (1) Water and air are incompressible Newtonian fluids. (2) The heat and mass transfer between the gas and liquid phases were not considered, and the two-phase flow was assumed to be an isothermal flow. (3) Both the top slag in the ladle and the gas phase in the vacuum chamber were not considered. The liquid surface at the top of both the ladle and the vacuum chamber were assumed to be flat. (4) The gas bubble shape was assumed to be spherical, and their expansion, coalescence and breaking up were not considered. (5) A two-way coupling between water and air bubbles was used.

    The liquid flow field was obtained by solving the continuity equation, momentum equation, and turbulence model equations, as follows [44]:

    Continuity equation

    \frac{{\partial {\rho _{\text{l}}}}}{{\partial t}} + \nabla \cdot \left( {{\rho _{\text{l}}}{\boldsymbol{u}}} \right) = 0 (4)

    Momentum equation

    \frac{{\partial \left( {{\rho _{\text{l}}}{\boldsymbol{u}}} \right)}}{{\partial t}} + \nabla \cdot \left( {{\rho _{\text{l}}}{\boldsymbol{uu}}} \right) = - \nabla P + \nabla \cdot \left[ {{\mu _{{\text{eff}}}}\left( {\nabla {\boldsymbol{u}} + {{\left( {\nabla {\boldsymbol{u}}} \right)}^{T}}} \right)} \right] + {\boldsymbol{F}} + {\rho _{\text{l}}}{\boldsymbol{g}} (5)

    where u is liquid velocity, m·s–1; μeff is effective viscosity, Pa·s, which is the sum of laminar viscosity (μl) and turbulence viscosity (μt); F is the momentum transfer term exerted by the discrete phase, N·m–3; g is the gravitational acceleration rate, m·s−2.

    Turbulent properties of the flow field were solved by using standard k–ε turbulence model. The turbulence viscosity is related to the turbulent kinetic energy k (m2·s–2) and the dissipation rate ε (m2·s–3) as follows [44]:

    {\mu _{\text{t}}} = {\rho _{\text{l}}}{C_{\text{μ }}}\frac{{{k^2}}}{\varepsilon } (6)

    where k and ε can be obtained by solving the following equations [44]:

    {\rho _{\text{l}}}\frac{{\partial k}}{{\partial t}} + {\rho _{\text{l}}}{\boldsymbol{u}} \cdot \nabla k = \nabla \cdot \left[ {\left( {{\mu _{\text{l}}} + \frac{{{\mu _{\text{t}}}}}{{{\sigma _{\text{k}}}}}} \right)\nabla k} \right] + {P_{\text{k}}} - {\rho _{\text{l}}}\varepsilon (7)
    {\rho _{\text{l}}}\frac{{\partial \varepsilon }}{{\partial t}} + {\rho _{\text{l}}}{\boldsymbol{u}} \cdot \nabla \varepsilon = \nabla \cdot \left[ {\left( {{\mu _{\text{l}}} + \frac{{{\mu _{\text{t}}}}}{{{\sigma _ {\varepsilon }}}}} \right)\nabla \varepsilon } \right] + {C_ {{\varepsilon }1}}\frac{\varepsilon }{k}{P_{\text{k}}} - {C_ {{\varepsilon }2}}{\rho _{\text{l}}}\frac{{{\varepsilon ^2}}}{k} (8)

    where Cμ, Cε1, Cε2, σk, and σε are the constants in the turbulence model, with the value of 0.09, 1.44, 1.92, 1.0, and 1.3, respectively [4445]. The production rate of turbulence kinetic energy Pk is expressed as follows [44]:

    {P_{\text{k}}} = {\mu _{\text{t}}}\left[ {\nabla {\boldsymbol{u}}:\left( {\nabla {\boldsymbol{u}} + {{\left( {\nabla {\boldsymbol{u}}} \right)}^T}} \right)} \right] (9)

    Air gas is injected into the liquid through the nozzles. The movement of air bubbles can be obtained by solving the following equation [46]:

    \frac{{{\text{d}}{{\boldsymbol{x}}_{\text{p}}}}}{{{\text{d}}t}} = {{\boldsymbol{v}}_{\text{b}}} (10)

    where xp is the bubble position and vb is the bubble velocity. The velocity of air bubbles can be computed by the following equation:

    \frac{{{{\rm{d}}( {{m_{{\rm b}}}{{\boldsymbol{v}}_{{\rm b}}}} )}}}{{{\text{d}}t}} = {\overrightarrow{F}}_{\text{ D}}+{\overrightarrow{F}}_{\text{ L}}+{\overrightarrow{F}}_{\text{ VM}}+{\overrightarrow{F}}_{\text{ P}}+{\overrightarrow{F}}_{\text{ G}}+{\overrightarrow{F}}_{\text{ B}} (11)

    where mb is the mass of bubbles, {\overrightarrow{F}}_{\text{ D}} is the drag force, {\overrightarrow{F}}_{\text{ L}} is the lift force, {\overrightarrow{F}}_{\text{ VM}} is the virtual mass force, {\overrightarrow{F}}_{\text{ P}} is the pressure gradient force, {\overrightarrow{F}}_{\text{ G}} and {\overrightarrow{F}}_{\text{ B}} are the gravity and buoyancy force, respectively. These forces are usually considered to investigate the circulation flow rate and mixing time in a RH refining process [10,47].

    The drag force plays a decisive role in the momentum transfer between the gas and liquid phase. It can be expressed by the following equation [46]:

    {\overrightarrow{F}}_{\text{ D}} = \frac{{{m_{\text{b}}}\left( {{\boldsymbol{u}} - {{\boldsymbol{v}}_{\text{b}}}} \right)}}{{{\tau _{\text{b}}}}} (12)

    where τb is the particle relaxation time, which can be calculated by the following equation [46]:

    {\tau _{\text{b}}} = \frac{{4{\rho _{\text{g}}}d_{\text{b}}^{\text{2}}}}{{3{\mu _{\text{l}}}{C_{\text{ D}}}R{e_{\text{b}}}}} (13)

    where db is bubble diameter, Reb is the Reynolds number of air bubbles, CD is the drag coefficient, where Schiller and Naumann’s [48] drag law was used to calculate this value as follows:

    {C_{\text{D}}} = \left\{ {\begin{array}{*{20}{c}} {\dfrac{{24}}{{R{e_{\text{b}}}}}\left( {1 + 0.15Re_{\text{b}}^{0.687}} \right)}&{\begin{array}{*{20}{c}} {}&{} \end{array}R{e_{\text{b}}}\; \text{≤}\; 1000} \\ {0.44}&{\begin{array}{*{20}{c}} {}&{} \end{array}R{e_{\text{b}}} > 1000} \end{array}} \right. (14)
    R{e_{\text{b}}} = \frac{{{\rho _{\text{l}}^{}}\left| {{\boldsymbol{u}} - {{\boldsymbol{v}}_{\text{b}}}} \right|{d_{\text{b}}}}}{{{\mu _{\text{l}}^{}}}} (15)

    The bubble diameter was calculated through the empirical formula derived by Sano et al. [49] as follows:

    {d}_{\text{b}}=0.091{\left(\frac{{\sigma }_{\text{gl}}}{{\rho }_{\text{l}}}\right)}^{0.5}{{\boldsymbol{v}}}_{\text{b}, 0}^{0.44} (16)

    where σgl is the surface tension between air and water, N·m–1; vb,0 is the initial bubble velocity at the entrance which can be calculated by the following equation:

    {{\boldsymbol{v}}}_{\text{b}, 0}=\frac{Q}{{A}_{\text{nozzle}}} (17)

    where Anozzle is the total area of all the blowing nozzles, m2.

    The Saffman lift force can be expressed by the following equation [46]:

    {\overrightarrow{F}}_{\text{ L}} = 6.46r_{\text{b}}^{\text{2}}{\boldsymbol{L}}_{\text{v}}\sqrt {{\mu _{\text{l}}}{\rho _{\text{l}}}\frac{{\left| {{\boldsymbol{u}} - {{\boldsymbol{v}}_{\text{b}}}} \right|}}{{\left| {{\boldsymbol{L}}_{\text{v}}} \right|}}} (18)

    where rb is the radius of the bubble, m; Lv is relative velocity gradient, which is computed by the following equation [46]:

    {\boldsymbol{L}}_{\rm{v}} = \left( {{\boldsymbol{u}} - {{\boldsymbol{v}}_{\text{b}}}} \right) \times \left[ {\nabla \times \left( {{\boldsymbol{u}} - {{\boldsymbol{v}}_{\text{b}}}} \right)} \right] (19)

    The virtual mass force is proportional to the relative acceleration of the discrete phase, which can be calculated by the following equation [46]:

    {\overrightarrow{F}}_{\text{ VM}} = \frac{{{\text{π }}d_{\text{b}}^3{\rho _{\text{l}}}}}{{12}}\frac{{{\text{d}}\left( {{\boldsymbol{u}} - {{\boldsymbol{v}}_{\text{b}}}} \right)}}{{{\text{d}}t}} (20)

    The pressure gradient force can be expressed by the following equation [46]:

    {\overrightarrow{F}}_{\text{ P}} = \frac{{{\text{π }}d_{\text{b}}^{\text{3}}{\rho _{\text{l}}}}}{6}\left( {\frac{{\partial {\boldsymbol{u}}}}{{\partial t}} + {\boldsymbol{u}} \cdot \nabla {\boldsymbol{u}}} \right) (21)

    The gravity and buoyancy forces are given by the following equation [46]:

    {\overrightarrow{F}}_{\text{ G}}+{\overrightarrow{F}}_{\text{ B}} = {m_{\text{b}}}{\boldsymbol{g}}\frac{{\left( {{\rho _{\text{g}}} - {\rho _{\text{l}}}} \right)}}{{{\rho _\text{g}}}} (22)

    In order to consider the effect of turbulent fluctuations on bubble motion, the DRW (Discrete random walk) model was used. The instantaneous velocity can be expressed by the following equation:

    {\boldsymbol{ u}} =\boldsymbol{ \bar u }+ \boldsymbol{\acute{u}} (23)

    where \bar {\boldsymbol{u}} is the averaged velocity of the continuous phase, {\boldsymbol{\acute{u}}} is velocity fluctuation component, which is expressed as follows:

    {\boldsymbol{\acute{u}}} = \zeta \sqrt {\frac{{2k}}{3}} (24)

    where ζ is a zero mean, which has a unit variance and normally distributed random number.

    Tracer transport was solved by using the following equation [44]:

    \frac{{\partial C}}{{\partial t}} + \nabla \times \left( {{\boldsymbol{u}}C} \right) = \nabla \times \left[ {{D_{{\text{eff}}}}\nabla C} \right] (25)

    where C is tracer concentration; Deff is effective diffusion coefficient, m2·s–1, which is the sum of molecule diffusion coefficient (D0) and turbulent diffusion coefficient (Dt). D0 value is 0.551 × 10–9 m2·s–1 for NaCl solution at 25°C [50]. Dt can be expressed as the following equation:

    {D_{\text{t}}} = \frac{{{\mu _{\text{t}}}}}{{\rho S {c_{\text{t}}}}} (26)

    where Sct is the turbulent Schmidt number, which is equal to 1 in this study.

    A constant velocity was used as the inlet condition for gas blowing. The vacuum chamber pressure is fixed at 97.7 kPa. A non-slip wall boundary condition was used. A reflect boundary condition was used for bubbles once they touch a wall, and they escaped from the calculation domain after they pass by the liquid surface in the vacuum chamber. It was considered to be a fully developed flow at 270 s after gas blowing, which can be seen from the velocity changes at monitoring positions in Fig. 4. Then, 200 mL tracer was released at the same location as that in water model experiment to investigate its transport and the bath mixing behavior.

    Figure  4.  Predicted velocity changes at monitoring points for different nozzle angles: (a) 0°; (b) 30°; (c) 45°.

    The numerical model was solved by the commercial software COMSOL Multiphysics 5.6. The mesh with 550,000 grid number was used, as shown in Fig. 5. A fine grid with the size of 1 mm was used at the gas blowing inlet area. The separated constant (Newton) nonlinear method in COMSOL Multiphysics was used to treat non-linear equations. The backward difference formula was firstly used to deal with the convection term. Then, the generalized-alpha method was further used to obtain the final solution. These two methods are all second order scheme. The initial time step was set to 0.0001 s, and the maximum step size was 0.02 s. The convergence criteria were as follows: the residuals of all dependent variables were smaller than 1 × 10–3.

    Figure  5.  Grid configuration of the RH model.

    Experimental measurements and numerical simulation were carried out to investigate the mixing time in the RH water model with a conventional nozzle. The mixing time measurements were repeated 10 times to obtain the mean value. Fig. 6 shows the comparison between the model predictions and the measured values. It can be seen from that the deviations at point 1, 2, and 3 were 7.3%, 1.3%, and 7.3%, respectively. Here, the mixing time was defined as the time when a dimensionless concentration was within 3% deviation from the averaged value. With a 5% deviation standard, the mixing time deviations at point 1, 2, and 3 were 15.3%, 4.0%, and 3.1%, respectively. Therefore, the model can generally catch the mixing behavior inside the bath. In the following, the 3% deviation criterion was used to evaluate the bath mixing time.

    Figure  6.  Comparison of predicted and measured concentration change: (a) point 1; (b) point 2; (c) point 3.

    Fig. 7 shows the mixing times at point 1, 2, and 3 under different nozzle blowing angles. It can be seen from that the mixing time at each monitoring point was decreased after either 30° or 45° rotated nozzle. Specifically, when the nozzle rotated from a conventional 0° to 30°, the mixing time at monitoring point 1, 2, and 3 was shortened by 16.2, 24.7, and 9.3 s, respectively, which corresponds to a decrease of 21.3%, 28.2%, and 12.3%. Therefore, a rotated nozzle can effectively reduce the mixing time. With the nozzle angle changed from 30° to 45°, the mixing time change is very small, as shown in Fig. 7(d).

    Figure  7.  Effect of nozzle angle on predicted concentration change and mixing time: (a) point 1; (b) point 2; (c) point 3; (d) mixing time at different points and nozzle angles.

    In order to have a good observation on the mass transfer behavior induced by different nozzle angles, tracer was added at the lower part of the up-snorkel (Z = 690 mm cross section as shown in Fig. 3). Fig. 8 shows the variation of tracer concentration with time in the snorkels and the vacuum chamber under different nozzle angles. Compared to a conventional nozzle angle with 0°, the high tracer concentration region seems to move faster towards the down-snorkel when both 30° and 40° rotated nozzles were used. Furthermore, the tracer concentration is more uniform on the cross section of the up-snorkel.

    Figure  8.  Movement of the high tracer concentration region in the snorkels and vacuum chamber for different nozzle angles and times: (a) 0°; (b) 30°; (c) 45°.

    The mixing time and the refining efficiency are related to the circulation flow rate, which can be calculated by using the following equation:

    R{\text{ = }}{\rho _{\text{l}}} \times \int_{{A_0}} {{u_i}{\text{d}}{A_i}} (27)

    where R is the circulation flow rate, kg·s–1; ui is the velocity perpendicular to the snorkel cross section, m·s–1; Ai is each grid area on the cross section, m2; A0 is the cross-section area, m2.

    The cross-section of the down-snorkel, located at 700 mm above the bottom of the ladle as shown in Fig. 3, was selected to calculate the circulation flow rate. Fig. 9 shows the calculated values of the circulation flow rate under different nozzle angles. It can be seen that the circulation flow rate increases from 9 to 10.3 kg·s–1 when the nozzle angle increases from 0° to 30°. With a further increase to 45°, the circulation flow rate increases by only 0.1 kg·s–1. Therefore, the rotated nozzle injection can increase the circulation flow rate during RH refining. This also explains the faster movement of the high tracer concentration region in the vacuum chamber as shown in Fig. 8.

    Figure  9.  Effect of nozzle angles on circulation flow rate.

    In order to better understand the reason for a high circulation flow rate and a low mixing time, flow phenomena under different nozzle conditions were studied. Fig. 10 shows the streamline of the liquid flow under different nozzle angles. It can be seen that the vertical movement of liquid was observed when a conventional nozzle was used, while a rotational movement was found for both 30° and 45°. This means that the gas momentum provides the force to produce a rotational liquid flow in the up-snorkel. Due to the symmetrical setup of the conventional nozzles, such phenomena were not produced. Fig. 11 shows the velocity distribution on the vertical middle plane of the whole system. As can be seen, the liquid flow velocity in the up-snorkel significantly increased and the velocity distribution became more uniform when 30° and 45° rotated nozzle were used. With a conventional 0°, the gas moves upwards along the wall of the up-snorkel. Therefore, its influence region on liquid is relatively smaller. In addition, with both 30° and 45° rotated nozzle, the liquid velocity in the down-snorkel was also increased. This can also explain the obtained high circulation flow rate in Fig. 9.

    Figure  10.  Streamline of liquid flow under different nozzle angles: (a) 0°; (b) 30°; (c) 45°.
    Figure  11.  Velocity distribution on the vertical middle plane of the RH reactor under different nozzle angles: (a) 0°; (b) 30°; (c) 45°.

    To describe the bubble behavior in the RH system intuitively, Fig. 12 shows the trajectories and residence time of the bubbles under different nozzle angles. When the nozzle is not rotated, air bubbles move vertically into the vacuum chamber; when the nozzle was rotated, gas bubbles obtain a rotational movement and move upwards under the action of buoyancy effect. This can be clearly seen from Fig. 12(b) and (c). Under the effect of the considered force in Eq. (11), the spirally rising bubbles lead to the molten steel generating a rotational flow as shown in Fig. 10. In addition, the rotational flow makes bubbles moving towards the up-snorkel center. The average residence time of bubbles can be calculated using the following equation:

    Figure  12.  Trajectories and residence time of the bubbles: (a) 0°; (b) 30°; (c) 45°.
    {\bar t_{\text{b}}}= \frac{{\displaystyle\sum\limits_{i = 1}^{i={n_{\text{b}}}} {{t_{{\text{b,}}i}}} }}{{{n_{\text{b}}}}} (28)

    where tb,i is the residence time of a bubble, \bar t_{\text{b}} is the average residence time of bubbles, and nb is the number of bubbles (with the value of 128 in this work). The average residence time is about 0.6 s in the conventional case. When the nozzle rotates 30° and 45°, the average residence time can reach around 0.8 s. In addition, the bubble trajectory length increases in the rotated nozzle cases. These should be beneficial for the refining process that happens at the bubble/steel interface.

    In order to clearly understand the flow characteristics in the up-snorkel, Fig. 13 shows the velocity distribution on different cross sections of the up-snorkel under different nozzle angles. For the conventional nozzle case (Fig. 13(a)), the high velocity region was located at the near wall region, and the velocity is very small with the value of about 0.1 m·s–1 in the center region. After the nozzle is horizontally rotated, the fluid in the up-snorkel has formed a good rotational flow. The tangential velocity distribution and its magnitude under different nozzle angles are similar. At the near wall region, the tangential velocity can reach 0.5 m·s–1 as shown in Fig. 13(d)–(f). For the vertical velocity, it can be seen from Fig. 13(g)–(i) that a low velocity region was located at the snorkel center when a conventional nozzle was used. The velocity characteristics on the cross section at a top location can be found from Fig. 13(j)–(r). Generally, it can be seen that a high velocity region is still located at the near wall region for the conventional nozzle case, while it moves towards the snorkel center for rotated nozzle cases. The tangential velocity in Fig. 13(n)–(o) becomes more uniform compared to that of Fig. 13(e)–(f). For the vertical velocity, the near wall region shows a low value in Fig. 13(p)–(r) compared to that of Fig. 13(g)–(i). This is attributed to the gas plume moving towards the center region during it passing through the up-snorkel. Therefore, the rotated nozzle design significantly changed the flow behavior in the up-snorkel.

    Figure  13.  Velocity distribution on different sections of the up-snorkel under different nozzle angles: (a–c) total velocity at Z = 890 mm; (d–f) tangential velocity at Z = 890 mm; (g–i) axial velocity at Z = 890 mm; (j–l) total velocity at Z = 990 mm; (m–o) tangential velocity at Z = 990 mm; (p–r) axial velocity at Z = 990 mm.

    Fig. 14 shows the distributions of the velocity and turbulent kinetic energy on the cross section of the vacuum chamber. It can be seen from Fig. 14(a)–(c) that a symmetrical distribution along the Line 1 was observed for the conventional nozzle case. However, the velocity did not show a symmetrical distribution for the rotated nozzle cases. This is due to the rotational direction of the flow in the up-snorkel. Fig. 14(d)–(f) shows the turbulent kinetic energy distribution on the cross section. Among the three cases, the highest turbulent kinetic energy was obtained when a conventional nozzle was used. With a rotated nozzle angle, a general uniform total turbulent kinetic energy distribution can be obtained, and the high turbulent kinetic energy region enlarged as shown in Fig. 14(d)–(f).

    Figure  14.  Predicted properties of the flow field in the vacuum chamber: (a–c) total velocity; (d–f) turbulent kinetic energy.

    Fig. 15 shows the velocity distribution on different cross sections of the ladle under different nozzle angles. With a conventional nozzle, the velocity distribution on the ladle cross section shows a symmetrical distribution along the horizontal central line as shown in Fig. 15. However, the distribution is no longer symmetrical in the rotated nozzle cases. Due to the rotational flow in the up-snorkel, the velocity in the vacuum chamber was changed, as shown Fig. 14(a)–(c). This directly affects the flow behavior inside the down-snorkel and ladle. In addition, the velocity inside the ladle was large when a rotated nozzle angle was used. This is especially true for the near-wall region, and the velocity can reach 0.24 m·s–1 on the cross section of Z = 200 mm, with an increase of 20%.

    Figure  15.  Distribution of velocity on different cross sections of the ladle under different nozzle angles: (a–c) Z = 400 mm; (d–f) Z = 200 mm.

    Fig. 16 shows the predicted turbulent kinetic energy distribution on the Z = 400 mm plane of the ladle under different nozzle angles. With the rotated nozzles, a larger turbulent kinetic energy was found in the ladle, the maximum magnitude increased from 0.008 (conventional case) to 0.011 m2·s–2 (rotated case) compared to the conventional nozzle case, resulting in a more efficient mixing inside the ladle as well as a reduced mixing time as predicted previously.

    Figure  16.  Distribution of turbulent kinetic energy on the ladle cross section under different nozzle angles: (a) 0°; (b) 30°; (c) 45°.

    Water model experiment and numerical simulation were carried out to investigate the effect of gas blowing nozzle angle on the mixing time, circulation flow rate, and fluid flow during RH refining process. The following conclusions were obtained:

    (1) The numerical model predictions generally agree well with the measured values, which illustrates its reliability to describe the multiphase flow process in RH refining.

    (2) With a rotated nozzle angle of 30° and 45°, the circulation flow rate increased by 14.4% and 15.6%, respectively, compared to a conventional nozzle. Furthermore, the mixing time at monitoring point 1, 2, and 3 were shortened by 21.3%, 28.2%, and 12.3%, respectively.

    (3) When 30° and 45° rotated nozzles were used, a rotational movement of liquid was obtained. The liquid flow velocity in the snorkel was significantly increased and the velocity distribution is more uniform in the up-snorkel. The velocity inside the ladle increased especially at the near-wall region.

    (4) The residence time of bubbles was increased by a rotated gas blowing nozzle, which was extended by around 33.3% for the nozzle angle of both 30° and 45° of 128 bubbles.

    This work was financially supported by the National Natural Science Foundation of China (No. 51704062) and the Fundamental Research Funds for the Central Universities, China (No. N2025019). Jiahao Wang wishes to thank the comments and discussions on the mathematical model from Dr. Xiaobin Zhou from Anhui University of Technology, China.

    On behalf of all authors, the corresponding author states that there is no conflict of interest.

  • [1]

    K. Nakanishi, J. Szekely, and C.W. Chang, Experimental and theoretical investigation of mixing phenomena in the RH-vacuum process, Ironmaking Steelmaking, 2(1975), No. 2, p. 115.

    [2]

    K. Shirabe and J. Szekely, A mathematical model of fluid flow and inclusion coalescence in the RH vacuum degassing system, ISIJ Int., 23(1983), No. 6, p. 465. DOI: 10.2355/isijinternational1966.23.465

    [3]

    R. Tsujino, J. Nakashima, M. Hirai, and I. Sawada, Numerical analysis of molten steel flow in ladle of RH process, ISIJ Int., 29(1989), No. 7, p. 589. DOI: 10.2355/isijinternational.29.589

    [4]

    Y. Kato, H. Nakato, T. Fujii, S. Ohmiya, and S. Takatori, Fluid flow in ladle and its effect on decarburization rate in RH degasser, ISIJ Int., 33(1993), No. 10, p. 1088. DOI: 10.2355/isijinternational.33.1088

    [5]

    F. Ahrenhold and W. Pluschkell, Mixing phenomena inside the ladle during RH decarburization of steel melts, Steel Res., 70(1999), No. 8-9, p. 314. DOI: 10.1002/srin.199905646

    [6]

    S.K. Ajmani, S.K. Dash, S. Chandra, and C. Bhanu, Mixing evaluation in the RH process using mathematical modelling, ISIJ Int., 44(2004), No. 1, p. 82. DOI: 10.2355/isijinternational.44.82

    [7]

    P.A. Kishan and S.K. Dash, Prediction of circulation flow rate in the RH degasser using discrete phase particle modeling, ISIJ Int., 49(2009), No. 4, p. 495. DOI: 10.2355/isijinternational.49.495

    [8]

    J.H. Wei and H.T. Hu, Mathematical modelling of molten steel flow process in a whole RH degasser during the vacuum circulation refining process: Mathematical model of the flow, Steel Res. Int., 77(2006), No. 1, p. 32. DOI: 10.1002/srin.200606127

    [9]

    J.H. Wei and H.T. Hu, Mathematical modelling of molten steel flow process in a whole RH degasser during the vacuum circulation refining process: Application of the model and results, Steel Res. Int., 77(2006), No. 2, p. 91. DOI: 10.1002/srin.200606359

    [10]

    H.T. Ling, F. Li, L.F. Zhang, and A.N. Conejo, Investigation on the effect of nozzle number on the recirculation rate and mixing time in the RH process using VOF+DPM model, Metall. Mater. Trans. B, 47(2016), No. 3, p. 1950. DOI: 10.1007/s11663-016-0669-y

    [11]

    G.J. Chen, S.P. He, and Y.G. Li, Investigation of the air-argon-steel-slag flow in an industrial RH reactor with VOF–DPM coupled model, Metall. Mater. Trans. B, 48(2017), No. 4, p. 2176. DOI: 10.1007/s11663-017-0992-y

    [12]

    B. Wang, B.H. Zhu, and B. Zhang, Numerical simulation of gas–steel–slag multiphase flow in the vacuum chamber of the RH degasser, JOM, 73(2021), No. 10, p. 2920. DOI: 10.1007/s11837-021-04816-6

    [13]

    G.J. Chen and S.P. He, Hydrodynamic modeling of two-phase flow in the industrial Ruhrstahl-Heraeus degasser: Effect of bubble expansion models, Metall. Mater. Trans. B, 53(2022), No. 1, p. 208. DOI: 10.1007/s11663-021-02357-6

    [14]

    P. Shao, S. Liu, and X. Miao, CFD–PBM simulation of bubble coalescence and breakup in top blown-rotary agitated reactor, J. Iron Steel Res. Int., 29(2022), No. 2, p. 223. DOI: 10.1007/s42243-021-00636-9

    [15]

    Y.G. Park, K.W. Yi, and S.B. Ahn, The effect of operating parameters and dimensions of the RH system on melt circulation using numerical calculations, ISIJ Int., 41(2001), No. 5, p. 403. DOI: 10.2355/isijinternational.41.403

    [16]

    R.K. Hanna, T. Jones, R. Blake, and M. Millman, Water modelling to aid improvement of degasser performance for production of ultralow carbon interstitial free steels, Ironmaking Steelmaking, 21(1994), p. 37.

    [17]

    C.W. Li, G.G. Cheng, X.H. Wang, G.S. Zhu, and A.M. Cui, Mathematical model of RH blow argon mode affecting: Decarburization rate in ultra-low carbon steel refining, J. Iron Steel Res. Int., 19(2012), No. 5, p. 23. DOI: 10.1016/S1006-706X(12)60095-8

    [18]

    L.C. Trindade, J.J.M. Peixoto, C.A. Silva, et al., Influence of obstruction at gas-injection nozzles (number and position) in RH degasser process, Metall. Mater. Trans. B, 50(2019), No. 1, p. 578. DOI: 10.1007/s11663-018-1474-6

    [19]

    D. Mukherjee, A.K. Shukla, and D.G. Senk, Cold model-based investigations to study the effects of operational and nonoperational parameters on the Ruhrstahl-Heraeus degassing process, Metall. Mater. Trans. B, 48(2017), No. 2, p. 763. DOI: 10.1007/s11663-016-0877-5

    [20]

    J.H. Wei and H.T. Hu, Mathematical modelling of molten steel flow in a whole degasser during RH refining process, Ironmaking Steelmaking, 32(2005), No. 5, p. 427. DOI: 10.1179/174328105X48133

    [21]

    X.G. Ai, Y.P. Bao, W. Jiang, J.H. Liu, P.H. Li, and T.Q. Li, Periodic flow characteristics during RH vacuum circulation refining, Int. J. Miner. Metall. Mater., 17(2010), No. 1, p. 17. DOI: 10.1007/s12613-010-0103-4

    [22]

    V. Seshadri and S.L.D.S. Costa, Cold model studies of RH degassing process, ISIJ Int., 26(1986), No. 2, p. 133. DOI: 10.2355/isijinternational1966.26.133

    [23]

    Y.H. Li, Y.P. Bao, R. Wang, L.F. Ma, and J.S. Liu, Modeling study on the flow patterns of gas–liquid flow for fast decarburization during the RH process, Int. J. Miner. Metall. Mater., 25(2018), No. 2, p. 153. DOI: 10.1007/s12613-018-1558-y

    [24]

    L. Lin, Y.P. Bao, F. Yue, L.Q. Zhang, and H.L. Ou, Physical model of fluid flow characteristics in RH–TOP vacuum refining process, Int. J. Miner. Metall. Mater., 19(2012), No. 6, p. 483. DOI: 10.1007/s12613-012-0584-4

    [25]

    Q. Wang, S.Y. Jia, F.S. Qi, et al., A CFD study on refractory wear in RH degassing process, ISIJ Int., 60(2020), No. 9, p. 1938. DOI: 10.2355/isijinternational.ISIJINT-2019-768

    [26]

    B. Li and F. Tsukihashi, Modeling of circulation flow in RH degassing vessel water designed for two- and multi-legs operations, ISIJ Int., 40(2000), No. 12, p. 1203. DOI: 10.2355/isijinternational.40.1203

    [27]

    F. Obata, R. Waka, K. Uehara, K. Ito, and Y. Kawata, Circulation characteristics of RH degassing vessel water model with multi-legs, Tetsu-to-Hagane, 86(2000), No. 4, p. 225. DOI: 10.2355/tetsutohagane1955.86.4_225

    [28]

    H. Ling, L. Zhang, and C. Liu, Effect of snorkel shape on the fluid flow during RH degassing process: Mathematical modelling, Ironmaking Steelmaking, 45(2018), No. 2, p. 145. DOI: 10.1080/03019233.2016.1248700

    [29]

    T. Kuwabara, K. Umezawa, K.J. Mori, and H. Watanabe, Investigation of decarburization behavior in RH-reactor and its operation improvement, ISIJ Int., 28(1988), No. 4, p. 305. DOI: 10.2355/isijinternational1966.28.305

    [30]

    Z.F. Ren, Z.G. Luo, F.X. Meng, et al., Physical simulation on molten steel flow characteristics of RH vacuum chamber with arched snorkels, Adv. Mater. Sci. Eng., 2020(2020), art. No. 6525096. DOI: 10.1155/2020/6525096

    [31]

    G.J. Chen and S.P. He, Mixing behavior in the RH degasser with bottom gas injection, Vacuum, 130(2016), p. 48. DOI: 10.1016/j.vacuum.2016.04.028

    [32]

    R.D. Wang, Y. Jin, and H. Cui, The flow behavior of molten steel in an RH degasser under different ladle bottom stirring processes, Metall. Mater. Trans. B, 53(2022), No. 1, p. 342. DOI: 10.1007/s11663-021-02371-8

    [33]

    D.Q. Geng, H. Lei, and J.C. He, Simulation on flow field and mixing phenomenon in RH degasser with ladle bottom blowing, Ironmaking Steelmaking, 39(2012), No. 6, p. 431. DOI: 10.1179/1743281211Y.0000000090

    [34]

    D.Q. Geng, J.X. Zheng, K. Wang, et al., Simulation on decarburization and inclusion removal process in the Ruhrstahl-Heraeus (RH) process with ladle bottom blowing, Metall. Mater. Trans. B, 46(2015), No. 3, p. 1484. DOI: 10.1007/s11663-015-0314-1

    [35]

    S.P. He, G.J. Chen, and C.J. Guo, Investigation of mixing and slag layer behaviours in the RH degasser with bottom gas injection by using the VOF–DPM coupled model, Ironmaking Steelmaking, 46(2019), No. 8, p. 771. DOI: 10.1080/03019233.2017.1410948

    [36]

    J.F. Dong, C. Feng, R. Zhu, G.S. Wei, J.J. Jiang, and S.Z. Chen, Simulation and application of Ruhrstahl-Heraeus (RH) reactor with bottom-blowing, Metall. Mater. Trans. B, 52(2021), No. 4, p. 2127. DOI: 10.1007/s11663-021-02174-x

    [37]

    B. Li, Y. Luan, F. Qi, and H. Huo, Experimental investigation on circulation flow in RH refining system with swirling flow, J. Northeast Univ., 26(2005), No. 8, p. 726.

    [38]

    B.K. Li and F. Tsukihashi, Effect of rotating magnetic field on two-phase flow in RH vacuum degassing vessel, ISIJ Int., 45(2005), No. 7, p. 972. DOI: 10.2355/isijinternational.45.972

    [39]

    D.Q. Geng, H. Lei, and J.C. He, Effect of traveling magnetic field on flow, mixing, decarburization and inclusion removal during RH refining process, ISIJ Int., 52(2012), No. 6, p. 1036. DOI: 10.2355/isijinternational.52.1036

    [40]

    F.S. Qi, B.K. Li, and F. Tsukihashi, Behaviour of argon gas bubbles in an electromagnetic driven swirling flow, Steel Res. Int., 78(2007), No. 5, p. 409. DOI: 10.1002/srin.200705912

    [41]

    L.Y. Wang, Z.Y. Liu, M. He, et al., Flow fields control for bubble refinement induced by electromagnetic fields, J. Iron Steel Res. Int., 29(2022), No. 4, p. 575. DOI: 10.1007/s42243-021-00627-w

    [42]

    W.M. Haynes, CRC Handbook of Chemistry and Physics, 93rd eds., CRC Press, Florida, 2012, p. 5.

    [43]

    J. Han, X.D. Wang, and D.C. Ba, Coordinated analysis of multiple factors of argon blowing parameters on the effect of circulation flow rate in RH vacuum refining process, Vacuum, 109(2014), p. 68. DOI: 10.1016/j.vacuum.2014.05.007

    [44]

    COMSOL, CFD Module User’s Guide, Release 5.3, SWE, 2017, p. 44.

    [45]

    B.E. Launder and D.B. Spalding, Lectures in Mathematical Models of Turbulence, Academic Press, New York, 1972.

    [46]

    COMSOL, Particle Tracing Module User’s Guide, Release 5.6, SWE, 2020, p. 267.

    [47]

    Y. Luo, C. Liu, Y. Ren, and L.F. Zhang, Modeling on the fluid flow and mixing phenomena in a RH steel degasser with oval down-leg snorkel, Steel Res. Int., 89(2018), No. 12, art. No. 1800048. DOI: 10.1002/srin.201800048

    [48]

    L. Schiller and Z.A. Naumann, A drag coefficient correlation, Zeitschrift Des Vereins Deutscher Ingenieure, 77(1935), p. 318.

    [49]

    M. Sano and K. Mori, Dynamics of bubble swarms in liquid metals, Trans. Iron Steel Inst. Jpn, 20(1980), No. 10, p. 668. DOI: 10.2355/isijinternational1966.20.668

    [50]

    A. Ghaffari and A. Rahbar-Kelishami, MD simulation and evaluation of the self-diffusion coefficients in aqueous NaCl solutions at different temperatures and concentrations, J. Mol. Liq., 187(2013), p. 238. DOI: 10.1016/j.molliq.2013.08.004

Relative Articles

王璟丞, 刘珍童, 陈威, 陈宏亮, 张立峰. 两流连铸中间包换包过程多相流动及钢液二次氧化的数值模拟研究 [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-024-2909-5

View details

Int. J. Miner. Metall. Mater., 2022, 29(1): 70-77.

PDF View details

赵洪亮, 王镜琪, 刘风琴, Hong Yong Sohn. 铜PS转炉内的流域分布和混合时间研究 [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-020-2196-8

View details

Monideepa Mukherjee, Sumit Tiwari, Basudev Bhattacharya. Evaluation of factors affecting the edge formability of two hot rolled multiphase steels [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-018-1563-1

View details

Hong-liang Zhao, Pan Yin, Li-feng Zhang, Sen Wang. Water model experiments of multiphase mixing in the top-blown smelting process of copper concentrate [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-016-1360-7

View details

Int. J. Miner. Metall. Mater., 2015, 22(1): 102-110.

PDF View details

Yong-gui Shi, Yue Hao, Dong Wang, Jin-cheng Zhang, Peng Zhang, Xue-fang Shi, Dang Han, Zheng Chai, Jing-dong Yan. Effects of the flow rate of hydrogen on the growth of graphene [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-015-1049-3

View details

Int. J. Miner. Metall. Mater., 2010, 17(6): 709-714.

PDF View details

Dian-qiao Geng, Hong Lei, Ji-cheng He. Optimization of mixing time in a ladle with dual plugs [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-010-0378-5

View details

Xin-gang Ai, Yan-ping Bao, Wei Jiang, Jian-hua Liu, Peng-huan Li, Tai-quan Li. Periodic flow characteristics during RH vacuum circulation refining [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-010-0103-4

View details

Int. J. Miner. Metall. Mater., 2007, 14(6): 490-494.

PDF View details

Mingmei Zhu, Guanghua Wen, Ping Tang, Yufeng Wang. Dissymmetric flow phenomenon in a multistrand tundish [J]. 矿物冶金与材料学报(英文版). DOI: 10.1016/S1005-8850(07)60115-7

View details

Int. J. Miner. Metall. Mater., 2002, 9(3): 161-165.

PDF View details

Jiling Mao, Yanhua Shen, Jiang Liu, Sheng Ling. Simulative analysis for deep seabed mining lifting systems [J]. 矿物冶金与材料学报(英文版).

View details

Int. J. Miner. Metall. Mater., 1994, 1(1-2): 97-103.

PDF View details

郭鸿志, 朱浩, 赵沛, 傅杰. 钢包精炼炉两相流和重要工艺参数 [J]. 矿物冶金与材料学报(英文版).

View details

Citing articles(16)

Fengsheng Qi, Nan Ye, Zhongqiu Liu, et al. Numerical Study on the Decarburization Processes in Vacuum-Refining Degasser Using the Population Balance Method. Metallurgical and Materials Transactions B, 2025, 56(2): 1363. 必应学术
Jintao Song, Chao Chen, Zhijie Guo, et al. Analysis of Flow Field, Temperature, and Inclusion Evolution in a 12-Strand Tundish During Ladle Changeover Process. Metallurgical and Materials Transactions B, 2025. 必应学术
Gujun Chen, Jiang Yang, Yugang Li, et al. Mathematical Simulation of Decarburization with CO2 Injection During RH Refining of Ultra-Low-Carbon Steel. Metals and Materials International, 2025, 31(1): 167. 必应学术
More >

/

返回文章
返回