Next Article in Journal
A Novel Probabilistic Optimal Power Flow Method to Handle Large Fluctuations of Stochastic Variables
Previous Article in Journal
New Power Quality Indices for the Assessment of Waveform Distortions from 0 to 150 kHz in Power Systems with Renewable Generation and Modern Non-Linear Loads
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Study of the Gas-Liquid Two-Phase Flow in a Self-Designed Mixer for a Ga-R113 MHD System

1
Jiangsu Province Key Laboratory of Aerospace Power System, Key Laboratory of Thermal Environment and Thermal Structure for Aero Engines of Ministry of Industry and Information Technology, College of Energy and Power Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
2
College of Astronautics, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
3
Department of Chemical Engineering, Purdue University, West Lafayette, IN 47907, USA
4
School of Energy and Mechanical Engineering, Nanjing Normal University, Nanjing 210042, China
*
Authors to whom correspondence should be addressed.
Energies 2017, 10(10), 1629; https://doi.org/10.3390/en10101629
Submission received: 29 September 2017 / Revised: 14 October 2017 / Accepted: 15 October 2017 / Published: 17 October 2017

Abstract

:
Liquid metal MHD (Magneto-Hydro-Dynamic) systems can be employed to produce electricity from a wide range of heat resources. In such a system, a low-boiling organic fluid and a high-temperature liquid metal fluid mix. The former evaporates, and carries the latter to flow through an MHD channel, where the electricity is generated. The mixing process and the gas-liquid flow characteristics will have a significant effect on the power generating efficiency. In the present work, trifluorotrichloroethane (R113) was chosen as the organic fluid, and gallium (Ga) as the liquid metal, respectively. Numerical study was subsequently carried out on the gas-liquid flow and heat transfer in a self-designed spherical mixer. The effects of the main factors, including the inlet velocities and inlet temperatures of Ga and R113, were separately determined, with suggested values or ranges discussed in detail.

1. Introduction

At present, the ever-increasing energy cost, shortage of fossil fuel resources, and environmental pollution have aroused the development of high-efficiency energy conversion technologies [1]. MHD power-generation technology extends Faraday’s law to conductive fluids [2], which can be classified into high-temperature plasma MHD systems and LMMHD (Liquid Metal MHD) systems, based on the sort of working medium.
In high-temperature plasma MHD systems, the working gas has to be ionized at a very high temperature, usually up to over 5000 K [3,4], so as to reach an appropriate electrical conductivity [5]. Although ionization seeds can be used to lower this temperature, it still exceeds 2000~2500 K [6,7]. In addition, there is another crucial technical problem to overcome—the slagging [8].
Compared to high-temperature plasma MHD systems, LMMHD systems have no such problems. There are typically two types of fluids in an LMMHD system: the thermodynamic fluid (low-boiling organic medium) and the power-generating fluid (conductive liquid metal). The thermal efficiency is theoretically close to the Carnot cycle in that the low-boiling organic medium is continuously heated by the high-temperature liquid metal [9]. More significantly, LMMHD systems can be utilized for a wide range of heat resources, including solar, geothermal, fossil fuels, nuclear, and chemical reactions, etc.
The schematic of a two-phase LMMHD system is shown in Figure 1. The liquid metal is first heated up in the heater. In the mixer, the two fluids mix and the low-boiling organic medium evaporates. Subsequently, the gas-liquid two-phase fluids flow through the MHD generator and electricity is created. Finally, the fluids are separated and pumped back to the mixer through the upper and lower loops, respectively.
LMMHD systems are drawing increasing attention on various topics, including flow and heat transfer [10,11,12], magnetic effects [13,14], power-generating characteristics [15], and bubble behaviours [16], etc. For example, Kakarantzas et al. [12] investigated the MHD liquid metal flow and heat transfer in vertical annuli under a horizontal magnetic field by direct numerical simulations. They found that the fluid motion increases with the aspect ratio and annular gap, and the highest spatially averaged heat transfer rates are obtained for aspect ratios equal to one. Bakalis and Hatzikonstantinou [14] studied the MHD flow of a liquid metal in a curved annular channel, so as to examine the effect of curvature and the magnetic field on the velocity distribution. Schwarz and Froehlich [16] numerically simulated the upward motion of a single bubble in liquid metal exposed to an external magnetic field. The results showed that for large bubbles, the rise velocity increases first and then decreases; on the other hand, for small bubbles, the rise velocity decreases when strengthening the magnetic field.
However, there are rather few studies focusing on the mixing process and two-phase flow characteristics inside the mixer which can significantly affect the follow-up power-generation process. In our previous research, with tin and trifluorotrichloroethane (R113) chosen as the liquid metal and low-boiling working medium, respectively, a preliminary two dimensional numerical study was carried out on the two-phase flow characteristics [2]. Further research on the effects of the physical properties of liquid metal indicates that, compared to tin, Gallium (Ga) has a more beneficial influence on the flow and heat transfer process due to its higher heat capacity and conductivity [17]. Therefore, in the present paper, we will advance the research on a self-designed spherical mixer for two-phase LMMHD systems, with Ga and R113 as the working media. Their physical properties are presented in Table 1 and Table 2. A three-dimensional numerical study will be carried out, aiming to determine the effects of the main impacting factors on the gas-liquid two-phase flow characteristics.

2. Modelling

2.1. Model Set-Up

A self-designed mixer modeled by Unigraphics (UG) is shown in Figure 2. High-temperature liquid Ga is initially stored inside this adiabatic spherical mixer. R113 and liquid Ga enter the mixer from the bottom and left inlets, respectively. Once R113 makes contact with liquid Ga, it is heated up and evaporates into gas. The continuous expansion of R113 gas increases the inner pressure, and pushes the two-phase fluids out of the mixer from the right outlet. This mixer features a simple structure, and provides an appropriate route and space for mixing, flow, and heat transfer, especially for such gas-liquid two-phase media. In this way, the flow velocity is accelerated to a much higher value at the outlet, which is of great benefit to the electrical power generating in the follow-up MHD generator.

2.2. Numerical Model

2.2.1. Governing Equations

The multiphase models in FLUENT can be classified into the VOF (Volume of Fluid) model, Mixture model, and Eulerian model. Among them, the Mixture model is relatively suitable for simulating the mixing process of R113 and Ga, in combination with a standard k-ε two-equation turbulent model. In addition, the SIMPLE scheme is applied to resolve the pressure-velocity coupling equation, and the second-order upwind difference scheme is used for discretization, with the convergence precision of 10−6 to obtain a satisfactory accuracy. The governing equations [18] are as follows.
(1)
Continuity Equation
t ( ρ m ) + ( ρ m v m ) = 0
(2)
Momentum Equation
t ( ρ m v m ) + ( ρ m v m v m ) = p + [ μ m ( v m + v m T ) ] + ρ m g + F + ( k = 1 n α k ρ k v dr , k v dr , k ) ;
where v m is the mass-averaged velocity k = 1 n α k ρ k v k ρ m , ρ m is the mixture density k = 1 n α k ρ k , n is the number of phases, F is the body force, αk is the volume fraction of phase k, μm is the viscosity of the mixture k = 1 n α k μ k , and v dr , k is the drift velocity of phase k.
(3)
Energy Equation
t k = 1 n ( α k ρ k E k ) + k = 1 n [ α k v k ( ρ k E k + p ) ] = ( k eff T ) + S ;
Ek is the energy of phase k; S′ is the volumetric heat sources; and keff is the effective thermal conductivity k = 1 n α k ( k k + k t ) , where kk is the thermal conductivity of phase k and kt is the turbulent thermal conductivity.
(4)
Volume fraction equation of the second phase
t ( α k ρ k ) + ( α k ρ k ν k ) = ( α k ρ k ν dr , k ) .
A simple and accurate model proposed by Lee [19] and widely acknowledged is employed in the present paper to calculate the evaporation process of R113, with the relevant source terms expressed as follows.
R113 source term:
S M = r α R ρ R ( T T b T b ) ;
R113 (gas) source term:
S M = r α R ρ R ( T T b T b ) ;
Energy source term:
S E = r α R ρ R ( T T b T b ) Δ H .
where SM and SE are the mass and energy source terms, respectively; αR and ρR are the volume fraction and density of R113; T and Tb are the mixture temperature and boiling point of R113; ΔH is the latent heat; and r is the phase change factor.

2.2.2. Boundary Conditions

The velocity-inlet and pressure-outlet boundary conditions are applied to the computational domain, where the outlet pressure is set as the standard atmosphere. The boundary conditions for the pipe wall are impermeable, non-slip, and adiabatic, with an inner roughness at 5 × 10−5 m. Gravity acceleration is taken into consideration, whereas the radiant heat exchange between the working media is ignored.

2.2.3. Independence Verification

The 3D computational domain with meshes is depicted in Figure 3. The mesh number has a great influence on the simulation accuracy. Generally, increasing the mesh number will, on the one hand, obtain more accurate simulation results; on the other hand, it will result in a much longer computing time. In order to balance the simulation accuracy and computing time, it is necessary to perform the verification of grid independence.
The mesh number is set as 2.1 × 105, 3.7 × 105, 6.0 × 105, 8.9 × 105, and 12.1 × 105, respectively, and the corresponding outlet velocity (vlmo) and volume fraction of liquid Ga (VFo) are plotted in Figure 4. It can be seen that the values fluctuate after the mesh number exceeds 6.0 × 105, which can be consequently recognized as the appropriate mesh number.
The numerical method in this paper is testified by simulating a gas-liquid flow process in the literature [20], as shown in Figure 5. The figure describes the axial pressure distribution at different Reynolds numbers, and the simulated results acquired by the present method are in good agreement with the experiments, with the maximum error of less than 10%.

3. Results and Discussion

3.1. Effects of vlmi

The volume fraction distribution of liquid Ga (VF) and the flow field across the longitudinal section (Z = 0) under different inlet velocities of liquid Ga (vlmi) are depicted in Figure 6. As vlmi increases, the falling position of liquid Ga gradually rises, and the evaporation area (or mixing area) enclosed by the high-temperature Ga enlarges, which is beneficial to R113 evaporation.
The controlling variables method is applied in the present paper. Keeping the inlet temperature of Ga (Tlmi), inlet velocity of R113 (vRi), and inlet temperature of R113 (TRi) constant at 573 K, 0.6 m·s−1, and 310 K, respectively, the inlet velocity of liquid Ga (vlmi) varies from 0.6 m·s−1 to 3.0 m·s−1. It can be noticed from Figure 7 that as vlmi increases, all of the outcomes rise, with the outlet velocity of Ga (vlmo) increasing from 3.46 m·s−1 to 7.81 m·s−1, outlet volume fraction of Ga (VFo) from 18.78% to 39.63%, and evaporation rate of R113 (ER) from 35.42% to 64.14%, respectively. It is understandable that more liquid Ga is provided due to the increasing inlet velocity. In addition, the evaporation area increases, as shown in Figure 6. They both facilitate the evaporating process and improve the carrying ability of R113. As a result, vlmo increases. Since more liquid Ga is supplied and conveyed, VFo increases accordingly.
Although in general, increasing vlmi has a positive effect on the outlet flow characteristics, it should be pointed out that there should be a limit on it. Figure 7a shows that the outlet temperature of liquid Ga (Tlmo) is very close to its inlet temperature (573 K) when the velocity exceeds 2.4 m·s−1, suggesting that the isothermal expansion of the two-phase mixture is realized and the optimal thermodynamic efficiency is achieved. A further increase of vlmi will not obviously improve the evaporation area, as shown in Figure 6d,e. Conversely, it will consume more pump power and reduce the overall efficiency of the system. Taking the above factors into consideration, the suggested range for vlmi is between 1.8 m·s−1 and 2.4 m·s−1.

3.2. Effects of Tlmi

Keeping the inlet velocity of liquid Ga (vlmi), inlet velocity of R113 (vRi), and inlet temperature of R113 (TRi) constant at 2.4 m·s−1, 0.6 m·s−1, and 310 K, respectively, the variations of inlet temperature of liquid Ga (Tlmi) are presented in Figure 8. This figure shows that with Tlmi increasing, the outlet velocity of Ga (vlmo) rises from 6.48 m·s−1 to 8.59 m·s−1, and the evaporation rate of R113 (ER) rises from 48.73% to 80.20%; whereas the outlet volume fraction of Ga (VFo) declines from 38.33% to 29.10%. The rising Tlmi is beneficial to R113 evaporation; thus, the carrying-ability of R113 gas is promoted and the outlet velocity of Ga (vlmo) rises. Nevertheless, the increasing volumetric proportion of R113 gas will adversely reduce VFo and the electrical conductivity of the fluids. The reason for keeping VFo above a certain value will be subsequently discussed in further detail.

3.3. Effects of vRi

Keeping the inlet velocity of liquid Ga (vlmi), inlet temperature of liquid Ga (Tlmi), and inlet temperature of R113 (TRi) constant at 2.4 m·s−1, 573 K, and 310 K, respectively, the inlet velocity of R113 (vRi) ranges from 0.2 m·s−1 to 1.0 m·s−1. A higher vRi results in a higher R113 supply. As the thermodynamic fluid, its carrying-ability is enhanced. As a result, a higher outlet velocity of Ga (vlmo) is obtained (Figure 9a) from 4.44 m·s−1 to 9.43 m·s−1. On the contrary, the outlet volume fraction of Ga (VFo) declines from 55.61% to 26.81% due to the higher volumetric percentage of R113 gas. Because of the increased average velocity, the heat-transfer time length between the two fluids is reduced; thus, the evaporation rate of R113 (ER) decreases from 77.85% to 51.65% (Figure 9b). Still, it should be pointed out that VFo should be controlled and this will be discussed in the next section.

3.4. Effects of TRi

Keeping the inlet velocity of liquid Ga (vlmi), inlet temperature of liquid Ga (Tlmi), and inlet velocity of R113 (vRi) constant at 2.4 m·s−1, 573 K, and 0.6 m·s−1, respectively, the variations of inlet temperature of R113 (TRi) are presented in Figure 10. The temperature difference between the two fluids is reduced with TRi, which raises ER, and lowers VFo slightly from 34.99% to 33.39% accordingly. Besides, since more R113 carrying gas is generated, the outlet velocity of Ga (vlmo) rises from 7.17 m·s−1 to 7.42 m·s−1. Note that a very low VFo may result in the R113 gas films covering the pipe walls in the follow-up MHD power-generating channel. In this situation, the annular flow pattern may be formed and leads to a loose contact between Ga and the electrodes. It is advised by Wallis [21] that VFo be higher than 20% so as to avoid the above problem.
Specifically, Figure 11 provides more detailed distributions for three different volume fractions of Ga along Y = 0 at the outlet cross section. Note that at VFo = 50% and VFo = 40%, the volume fraction of liquid Ga (VF) in the vicinity of the pipe wall is much higher than those at other locations, which is beneficial to power generation. However, with VFo decreasing to 30%, the differences of VF among different locations are reduced, with the VF adjacent to the pipe wall decreasing dramatically from 67% to 37%. This indicates that the proportion of continuous R113 gas increases, which will have a detrimental impact on the power generating process by detaching the liquid metal from the electrodes.

4. Conclusions

Liquid metal MHD systems can be utilized to produce electricity from varieties of heat resources. The mixing process and the gas-liquid flow characteristics will have a significant impact on the power-generating efficiency. In the present work, a three-dimensional numerical study is conducted on the above process in a self-designed spherical mixer, with liquid metal gallium (Ga) selected as the working medium. The impacts of the main parameters are determined, respectively, with the conclusions as follows:
(1)
The evaporation area enlarges with the inlet velocity of liquid Ga (vlmi), which is beneficial to R113 evaporation.
(2)
Generally, increasing vlmi plays a positive role in the gas-liquid flow characteristics, because vlmo, VFo and ER increase with vlmi. However, an excessively higher vlmi will result in an overload on the pump power, and consequently reduce the whole efficiency. The suggested range for vlmi is 1.8 m·s−1 to 2.4 m·s−1.
(3)
The inlet temperatures of liquid Ga (Tlmi) and R113 (TRi) have similar impacts on the gas-liquid mixing and flow characteristics. With Tlmi or TRi increasing, vlmo, Tlmo, and ER rise, while VFo declines. As the thermodynamic fluid, a higher inlet velocity of R113 (vRi) will, on one hand, obtain a higher vlmo; on the other hand, it will reduce Tlmo, ER, and VFo.
(4)
It is suggested that VFo be kept above a certain value. Otherwise an undesirable annular flow pattern may be formed, which has a detrimental impact on the power-generating process by detaching the liquid metal from the electrodes.
(5)
It is advised that future research work be centered on experimental investigations, microscopic bubble motions, economic performance, and commercialized operations in this field.

Acknowledgments

This work is supported by the “National Natural Science Foundation of China” (No. 11675077; 51506087); “Fundamental Research Funds” (No. JCKY2013203B003); “China Postdoctoral Science Foundation” (No. 2015M571747); “China Scholarship Council” (No. 201606835064); “the Fundamental Research Funds for the Central Universities” (No. NJ20160041); and the “Foundation of Graduate Innovation Center in Nanjing University of Aeronautics and Astronautics (NUAA)” (No. kfjj20160210).

Author Contributions

Peng Lu took direct charge of this research. Hulin Huang provided ideas and instructions. Xingwen Zheng, Lulu Fang, Shu Xu, and Yezhen Yu participated in this research.

Conflicts of Interest

There are no conflicts of interest to declare.

Nomenclature

Symbols
EenergyJ·kg−1
ERevaporation rate%
F body forceN
g gravitational accelerationm·s−2
kthermal conductivityW·m−1·K−1
nnumber of phases
ppressurePa
Rradiusm
ReReynolds number
rthe factor of phase changes−1
Ssource termkg·m3·s−1 or kJ·m3·s−1
Svolumetric heat sourcesW·m−3
ttimes
TtemperatureK
vvelocitym·s−1
v mass-average velocitym·s−1
VFvolume fraction of liquid metal%
Xaxis Xm
Yaxis Ym
Zaxis Zm
Greek symbols
αvolume fraction%
ΔHlatent heatkJ·kg−1
ρdensitykg·m−3
μviscositykg·m−1·s−1
Subscripts
bboiling point
drdrift
Eenergy
effeffective
iinlet
kphase k
lliquid
lmliquid metal
Mmass
mmixture
ooutlet
RR113
tturbulence model

References

  1. Koltsaklis, N.E.; Kopanos, G.M.; Georgiadis, M.C. Design and operational planning of energy networks based on combined heat and power units. Ind. Eng. Chem. Res. 2014, 53, 16905–16923. [Google Scholar] [CrossRef]
  2. Lu, P.; Zheng, X.Y.; Yang, P.J.; Fang, L.L.; Huang, H.L. Numerical investigation into the vapor-liquid flow in the mixer of a liquid metal magneto-hydro-dynamic system. RSC Adv. 2017, 7, 35765–35770. [Google Scholar] [CrossRef]
  3. Sakai, T.; Matsumoto, M.; Murakami, T.; Okuno, Y. Numerical simulation of power generation characteristics of a disk MHD generator with high-temperature inert gas plasma. Electr. Eng. Jpn. 2012, 179, 23–30. [Google Scholar] [CrossRef]
  4. Tanaka, M.; Aoki, Y.; Zhao, L.; Okuno, Y. Experiments on high-temperature xenon plasma magnetohydrodynamic power generation. IEEE Trans. Plasma Sci. 2016, 44, 1241–1246. [Google Scholar] [CrossRef]
  5. Komatsu, F.; Tanaka, M.; Murakami, T.; Okuno, Y. Experiments on high-temperature inert gas plasma MHD electrical power generation with hall and diagonal connections. Electr. Eng. Jpn. 2015, 193, 17–23. [Google Scholar] [CrossRef]
  6. Li, Y.W.; Zhang, B.L.; Gao, L.; Duan, C.D.; Zhang, L.; Wang, Y.T.; He, G.Q. Research on generation and control of high temperature gas in MHD power generation. J. Propuls. Technol. 2017, 38, 1419–1426. [Google Scholar]
  7. Tanaka, M.; Murakami, T.; Okuno, Y. Plasma characteristics and performance of magnetohydrodynamic generator with high-temperature inert gas plasma. IEEE Trans. Plasma Sci. 2014, 42, 4020–4025. [Google Scholar] [CrossRef]
  8. Kayukawa, N. Open-cycle magnetohydrodynamic electrical power generation: A review and future perspectives. Prog. Energy Combust. Sci. 2004, 30, 33–60. [Google Scholar] [CrossRef]
  9. Wu, Q.; Schubring, D.L.; Sienicki, J.J. Feasibility analysis of two-phase MHD energy conversion for liquid metal cooled reactors. Nucl. Eng. Des. 2007, 237, 2114–2119. [Google Scholar] [CrossRef]
  10. Prinz, S.; Bandaru, V.; Kolesnikov, Y.; Krasnov, D.; Boeck, T. Numerical simulations of magnetohydrodynamic flows driven by a moving permanent magnet. Phys. Rev. Fluids 2016, 1. [Google Scholar] [CrossRef]
  11. Sivakumar, R.; Vimala, S.; Sekhar, T.V.S. Influence of induced magnetic field on thermal MHD flow. Numer. Heat Transf. A 2015, 68, 797–811. [Google Scholar] [CrossRef]
  12. Kakarantzas, S.C.; Benos, L.T.; Sarris, I.E.; Knaepen, B.; Grecos, A.P.; Vlachos, N.S. MHD liquid metal flow and heat transfer between vertical coaxial cylinders under horizontal magnetic field. Int. J. Heat Fluid Flow 2017, 65, 342–351. [Google Scholar] [CrossRef]
  13. Rizzo-Sierra, J.A.; Lopez-Hernandez, O.I. Numerical study on the magnetohydrodynamics of an oscillatory flow under inductionless and core-side-layer approximations. Rev. Mex. Fis. 2016, 62, 369–380. [Google Scholar]
  14. Bakalis, P.A.; Hatzikonstantinou, P.M. Effect of curvature and magnetic field on MHD flow of a liquid metal in a curved annular duct. Int. J. Numer. Methods Heat Fluid Flow 2015, 25, 1818–1833. [Google Scholar] [CrossRef]
  15. Takahashi, R.; Matsuo, M.; Ono, M.; Harii, K.; Chudo, H.; Okayasu, S.; Ieda, J.; Takahashi, S.; Maekawa, S.; Saitoh, E. Spin hydrodynamic generation. Nat. Phys. 2016, 12, 52–56. [Google Scholar] [CrossRef]
  16. Schwarz, S.; Froehlich, J. Numerical study of single bubble motion in liquid metal exposed to a longitudinal magnetic field. Int. J. Multiph. Flow 2014, 62, 134–151. [Google Scholar] [CrossRef]
  17. Li, W.; Lu, P.; Zheng, X.W.; Huang, H.L. Numerical simulation of working fluid effects on two-phase flow in lmmhd power generation system. J. Jiangsu Univ. (Nat. Sci. Ed.) 2017, 38, 434–439. [Google Scholar]
  18. Fluent Inc. Fluent User’s Guild; Fluent Inc.: Lebanon, NH, USA, 2006. [Google Scholar]
  19. Lee, W.H. A Pressure Iteration Scheme for Two-Phase Flow Modeling; Hemisphere: Washington, DC, USA, 1980. [Google Scholar]
  20. Ahmadpour, A.; Noori Rahim Abadi, S.M.A.; Kouhikamali, R. Numerical simulation of two-phase gas—Liquid flow through gradual expansions/contractions. Int. J. Multiph. Flow 2016, 79, 31–49. [Google Scholar] [CrossRef]
  21. Wallis, G.B. One-Dimensional Two-Phase Flow; McGraw-Hill: New York, NY, USA, 1969. [Google Scholar]
Figure 1. Schematic of a two-phase LMMHD system.
Figure 1. Schematic of a two-phase LMMHD system.
Energies 10 01629 g001
Figure 2. The self-designed mixer model built by UG.
Figure 2. The self-designed mixer model built by UG.
Energies 10 01629 g002
Figure 3. The 3D computational domain with meshes.
Figure 3. The 3D computational domain with meshes.
Energies 10 01629 g003
Figure 4. Independent verification of grid number.
Figure 4. Independent verification of grid number.
Energies 10 01629 g004
Figure 5. Comparison of the simulated results with the experimental data in the literature.
Figure 5. Comparison of the simulated results with the experimental data in the literature.
Energies 10 01629 g005
Figure 6. VF distribution and flow field across the longitudinal section (Z = 0) under different vlmi: (a) vlmi = 0.6 m·s−1; (b) vlmi = 1.2 m·s−1; (c) vlmi = 1.8 m·s−1; (d) vlmi = 2.4 m·s−1; (e) vlmi = 3.0 m·s−1.
Figure 6. VF distribution and flow field across the longitudinal section (Z = 0) under different vlmi: (a) vlmi = 0.6 m·s−1; (b) vlmi = 1.2 m·s−1; (c) vlmi = 1.8 m·s−1; (d) vlmi = 2.4 m·s−1; (e) vlmi = 3.0 m·s−1.
Energies 10 01629 g006
Figure 7. Impacts of vlmi on the gas-liquid two-phase flow characteristics: (a) Impacts of vlmi on the vlmo and Tlmo; (b) Impacts of vlmi on the VFo and ER.
Figure 7. Impacts of vlmi on the gas-liquid two-phase flow characteristics: (a) Impacts of vlmi on the vlmo and Tlmo; (b) Impacts of vlmi on the VFo and ER.
Energies 10 01629 g007
Figure 8. Impacts of Tlmi on the gas-liquid two-phase flow characteristics: (a) Impacts of Tlmi on the vlmo and Tlmo; (b) Impacts of Tlmi on the VFo and ER.
Figure 8. Impacts of Tlmi on the gas-liquid two-phase flow characteristics: (a) Impacts of Tlmi on the vlmo and Tlmo; (b) Impacts of Tlmi on the VFo and ER.
Energies 10 01629 g008
Figure 9. Impacts of vRi on the gas-liquid two-phase flow characteristics: (a) Impacts of vRi on the vlmo and Tlmo; (b) Impacts of vRi on the VFo and ER.
Figure 9. Impacts of vRi on the gas-liquid two-phase flow characteristics: (a) Impacts of vRi on the vlmo and Tlmo; (b) Impacts of vRi on the VFo and ER.
Energies 10 01629 g009aEnergies 10 01629 g009b
Figure 10. Impacts of TRi on the gas-liquid two-phase flow characteristics: (a) Impacts of TRi on the vlmo and Tlmo; (b) Impacts of TRi on the VFo and ER.
Figure 10. Impacts of TRi on the gas-liquid two-phase flow characteristics: (a) Impacts of TRi on the vlmo and Tlmo; (b) Impacts of TRi on the VFo and ER.
Energies 10 01629 g010aEnergies 10 01629 g010b
Figure 11. VF distributions along Y = 0 at the outlet cross section.
Figure 11. VF distributions along Y = 0 at the outlet cross section.
Energies 10 01629 g011
Table 1. Physical properties of Ga.
Table 1. Physical properties of Ga.
Physical PropertiesLiquid Gallium
Molar mass (g·mol−1)69.723
Density (kg·m−3)5904
Melting point (K)303
Heat capacity (J·kg−1·K−1)383.52
Heat conductivity (W·m−1·K−1)58
Viscosity (kg·m−1·s−1)1.94 × 10−3
Table 2. Physical properties of R113 and R113 gas (R113 g).
Table 2. Physical properties of R113 and R113 gas (R113 g).
Physical PropertiesR113R113 g
Molar mass (g·mol−1)187.376187.376
Density (kg·m−3)15657.38
Boiling point (K)321
Latent heat of vaporization (kJ·kg−1)146.7
Heat capacity (J·kg−1·K−1)912673
Heat conductivity (W·m−1·K−1)0.06570.0778
Viscosity (kg·m−1·s−1)4.97 × 10−41.08 × 10−5
Standard-state enthalpy (J·kmol−1)−8 × 108−6.95 × 108

Share and Cite

MDPI and ACS Style

Lu, P.; Zheng, X.; Fang, L.; Huang, H.; Xu, S.; Yu, Y. Numerical Study of the Gas-Liquid Two-Phase Flow in a Self-Designed Mixer for a Ga-R113 MHD System. Energies 2017, 10, 1629. https://doi.org/10.3390/en10101629

AMA Style

Lu P, Zheng X, Fang L, Huang H, Xu S, Yu Y. Numerical Study of the Gas-Liquid Two-Phase Flow in a Self-Designed Mixer for a Ga-R113 MHD System. Energies. 2017; 10(10):1629. https://doi.org/10.3390/en10101629

Chicago/Turabian Style

Lu, Peng, Xingwen Zheng, Lulu Fang, Hulin Huang, Shu Xu, and Yezhen Yu. 2017. "Numerical Study of the Gas-Liquid Two-Phase Flow in a Self-Designed Mixer for a Ga-R113 MHD System" Energies 10, no. 10: 1629. https://doi.org/10.3390/en10101629

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

Article Metrics

Back to TopTop