Next Article in Journal
Employee Financial Wellness Programs (EFWPs) as an Innovation in Incentive Systems of Energy Sector Enterprises in Poland during the COVID-19 Pandemic—Current Status and Development Prospects
Previous Article in Journal
A Conceptual Transdisciplinary Framework to Overcome Energy Efficiency Barriers in Ship Operation Cycles to Meet IMO’s Initial Green House Gas Strategy Goals: Case Study for an Iranian Shipping Company
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Study on Flow and Heat Transfer Characteristics of Supercritical CO2 in Zigzag Microchannels

1
School of Mechanical Engineering, Hunan University of Arts and Science, Changde 415000, China
2
School of Aeronautic Science and Engineering, Beihang University, Beijing 100191, China
*
Author to whom correspondence should be addressed.
Energies 2022, 15(6), 2099; https://doi.org/10.3390/en15062099
Submission received: 10 February 2022 / Revised: 9 March 2022 / Accepted: 11 March 2022 / Published: 13 March 2022

Abstract

:
The zigzag channel is the uppermost channel type of an industrial printed circuit heat exchanger (PCHE). The effect of geometric properties on the flow and heat transfer performance of the channel is significant to the PCHE design and optimization. Numerical investigations were conducted on the flow and heat transfer characteristics of supercritical CO2 (sCO2) in semicircular zigzag channels by computational fluid dynamics method. The shear stress transfer (SST) k–ω model was used as turbulence model and the National Institute of Standards and Technology (NIST) real gas model with REFPROP database was used to evaluate the thermophysical parameters of sCO2 in this numerical method. The effectiveness of the simulation method is verified by experimental data. Thermal hydraulic performance for zigzag channels with different pitch lengths, bending angles, and hydraulic diameters are studied comparatively based on this numerical method, with the boundary conditions which cover the pseudocritical point. The comparison results show that reducing the bending angle and pitch length will strengthen the effect of boundary layer separation on the leeward side of the wall and enhance the heat transfer performance, but the pressure drop of the channel will also increase, and the decrease of channel hydraulic diameter is beneficial to the heat transfer enhancement, but it is not as significant as that of the straight channel.

1. Introduction

A printed circuit heat exchanger (PCHE) is a type of compact heat exchanger with high efficiency, high application pressure, and high application temperature. It was invented in Australia in 1980 and promoted for commercial application by Heatric. It has broad application prospects in the fields of ultra-high-temperature gas-cooled reactors, floating liquefied natural gas storage units, and other industrial energy [1]. PCHE typically employs diffusion-bonded arrays of plates where microchannels are formed by chemical etching [2]. The typical cross section shape is semicircular and the hydraulic diameter in a PCHE passage is between 700 μm and 1.5 mm [3]. The flow channel geometries can be designed as straight, zigzag, S-shape, and airfoil-finned channels [4]. However, the zigzag-type channel is more widely used in industrial applications. CO2 is a nontoxic and inexpensive gas. It has excellent thermophysical properties (high specific heat, high thermal conductivity, and low viscosity) near the pseudocritical point, as shown in Figure 1, which can considerably enhance the heat transfer without sacrificing the hydraulic performance [5]. Consequently, the application of CO2 in PCHE has become the focus of researchers.
Heat transfer and hydraulic characteristics are the basis of PCHE thermal design. Various experimental and numerical investigations have been performed to optimize the channel structures, fluid mediums, and operation conditions. Nikitin and his team first published the experimental results of the flow and heat transfer characteristics of sCO2 in zigzag PCHE in 2016 and developed the correlations of Nu and f, while the correlations are only applicable to channels of the specific geometric parameters and Reynolds number range of those used in the experimental study [6]. Kruizenga et al. investigated the thermal–hydraulic performance of sCO2 in a straight channel with semicircular cross section using both experimental and numerical methods, the analysis results showed that the commercial computational fluid dynamics (CFD) software can well predict the internal heat transfer characteristics of the PCHE channel [7]. Saeed and Kim also conducted the numerical analysis of an sCO2 PCHE using ANSYS-CFX and validated the simulation results using published experimental data [8]. Tu and Zeng studied the flow and heat transfer performance in semicircular straight channels of sCO2 fluid for both cooling and heating process using CFD method. A modified model based on Douglas A. Olson [9] correlation was proposed to predict the heat transfer performance of sCO2 in semicircular channels for both heating and cooling conditions [10]. These literature conclusions fully confirm the feasibility of the numerical method.
Subsequently, a large number of studies focused on finding better channel types. Kim et al. compared the heat transfer and hydraulic performance of sCO2 in PCHE with zigzag and airfoil-shaped fins. It was found that the thermal performance of the airfoil fin was close to that of zigzag, but the airfoil fin had lower pressure loss [11]. Mohammed et al. investigated the effect of channel shapes (zigzag, curve, and step) on the thermal and hydraulic performance of PCHE and found that the zigzag channels have the highest value of heat transfer coefficient and pressure drop [12]. Matsuo et al. conducted numerical simulations of three different channel types (zigzag, chamfered zigzag, airfoil) to study the geometrical effects on the local heat transfer coefficient and pressure drop for supercritical CO2 in PCHE and developed new correlations for Nu of the zigzag channels [13]. In the research of [14,15,16], the PCHE with zigzag channel and discontinuous S-shaped fins were numerically and experimentally investigated, and it was found that the S-shaped channel can significantly improve the hydraulic performance while keeping almost equal heat transfer performance compared to the zigzag channels.
There are also studies focusing on thermal hydraulic characteristics of the other fluid media in PCHE channels. Dai et al. studied the flow and heat transfer performance of water in the semicircular zigzag passage experimentally [17]. Minghui Chen et al. investigated the thermal–hydraulic performance of a zigzag channel PCHE using helium as fluid media [18].
Most of the studies on the flow and heat transfer characteristics of PCHE channels use a specific geometric parameter channel or straight channel, or the comparison between zigzag channel and other types of channels such as S-type and discontinuous airfoil fin type. The experimental data and empirical correlations given in these studies are also limited to a certain type of channel, and the operating pressure, fluid bulk temperature, and Reynolds number are limited to a certain range. However, there are few studies on the effect of geometric parameters on the flow and heat transfer performance of a zigzag channel. This paper aims at modeling the forced convection heat transfer of CO2 within the zigzag channels, which are the main channel type of PCHE, and studies the effects of its main geometric parameters (hydraulic diameter, pitch length, and bending angle) on its internal flow and heat transfer parameters, especially near the pseudocritical point. The numerical method and analysis results of this study can be used as a reference for PCHE industrial design and channel performance investigation.

2. Numerical Modeling

A numerical method for analyzing the steady-state flow and heat transfer properties of a zigzag channel is defined. In this method, ANSYS Fluent 2019 is used to solve the governing equations of the steady turbulent flow of sCO2 in the zigzag channels. The NIST real gas model with REFPROP V9.1 database was used to evaluate the thermodynamics and transport of approximately of CO2. Yoon et al. [19] and Ren et al. [20] conducted comparative studies using STD k–e, realizable k–e, and SST (shear stress transport) k–ω turbulence models to simulate the thermal–hydraulic performance of sCO2 intube-flowing and found that the SST k–ω turbulence model gives the best quantitative prediction. The same conclusion was likewise reached in [21,22,23,24]. Therefore, the SST k–ω model is adopted for further analysis in this study. The pressure-based coupled algorithm was used to establish the coupling of velocity and pressure. The numerical simulation is considered convergent as all iterative residuals of the governing equations are less than 10−5, and area-weighted average outlet temperature and area-weighted average inlet pressure are stable.

2.1. Geometry and Boundary Conditions

The flow and heat transfer performance of CO2 in horizontal zigzag channels is investigated in this paper. Figure 2 shows a schematic diagram of the computational domain with the boundary condition adopted in this research. The study will consider the effects of different geometric factors, including hydraulic diameter, bending angle, and pitch length on the internal flow and heat transfer characteristics of the channel, as shown in Table 1. To check the independent effect of the geometric parameter, the comparative study only changes one geometric parameter at a time, and the rest remains unchanged. For example, for investigating the effect of bend angle θ on the channel flow and heat transfer performance, four types of bend angle, 100°, 115°, 140°, and 180° (straight), are considered for the geometric model, while Lp is fixed to 7.75 mm and dh is fixed to 2 mm.
As shown in Table 2, the inlet temperature was changed from 280 K to 350 K to ensure the bulk temperature Tb covers the pseudocritical point of CO2 for this analysis. The outlet pressure of CO2 varies between 7.5 MPa and 9 MPa to keep its condition near the pseudocritical point. Constant wall heat flux (12 kW/m2 for heating case and −12 kW/m2 for cooling case) and mass flux 200 kg/m2-s were adopted in this numerical simulation. The thermal properties of CO2 were obtained from NIST Database (REFPROP V9.1).

2.2. Governing Equations

The governing equations regarding the continuity, momentum, and energy are expressed as Equations (1)–(3).
(1) Continuity equation:
x i ( ρ u i ) = 0
(2) Momentum equation [25]:
x j ( ρ u i u j ) = p x i + τ i j x j + ρ g i
(3) Energy equation:
x i ( u i (   ρ E + p ) ) = x i ( k e f f T x i )
where ρ g i is the gravitational body force, u i is overall velocity vector, T is the temperature, ρ is the density, u i is the velocity vector, E is the total energy, p is the static pressure, τ i j is the stress tensor, and k e f f is the effective conductivity ( k e f f = k + k t , k t is the turbulent thermal conductivity).
(4) Direct numerical simulation (DNS) of Navier–Stokes equations is the most accurate method for turbulence; however, it is not feasible in most situations to resolve the wide range of scales in time and space as the CPU requirements by far exceed the existing capacity. For this reason, averaging procedures such as the Reynolds method have to be applied to the Navier–Stokes equations to filter out the turbulent spectrum [26]. However, the averaging process introduces additional unknown terms into the transport equations that need to be provided by suitable turbulence models. The SST kω turbulence model is used in this paper and is described as follows.
The transport equations of the kω model are expressed as Equations (4) and (5):
x i ( ρ k u i ) = x j ( Γ k k x j ) + G k ρ β * k ω
x i ( ρ ω u i ) = x j ( Γ w ω x j ) + G ω ρ β ω 2
where G k is the generation of turbulence kinetic energy k due to mean velocity gradients, G ω is the generation of ω , and Γ k and Γ w represent the effective diffusivity of k and ω calculated by Equations (6) and (7).
Γ k = μ + μ t ( F 1 σ k 1 + 1 F 1 σ k 2 )
Γ ω = μ + μ t ( F 1 σ ω 1 + 1 F 1 σ ω 2 )
where σ k and σ ω are turbulent Prandtl numbers for k and ω , respectively, μ is the dynamic viscosity of the fluid, and F 1 is calculated as Equations (8)–(10).
F 1 = tan h ( ϕ 1 4 )
ϕ 1 = min [ max ( k 0.09 ω y , 500 μ ρ y 2 ω ) ,   4 ρ k σ ω 2 D ω y 2 ]
D ω = 2 ( 1 F 1 ) ρ σ ω 2 1 ω k x j ω x j
The turbulent viscosity μ t of the SST k ω model is calculated using Equations (11)–(15).
μ t = ρ k ω 1 max ( 1 a * , 2 S i j F 2 a 1 ω )
S i j = 1 2 ( u i x j + u j x i )
F 2 = tanh ( ϕ 2 2 )
ϕ 2 = max ( 2 k 0.09 ω y , 500 μ ρ y 2 ω )
a * = a * ( a 0 * + ρ k 6 μ ω 1 + ρ k 6 μ ω )
where μ t is the turbulent viscosity, y is the wall distance, and the constants used in the SST model are as follows: a * = 1 ,   a 1 = 0.31, σ k 1 = 1.176 , σ k 2 = 1 , σ ω 1 = 2 , σ ω 2 = 1.168 ,   a 0 * = 0.024 .
The second-order upwind scheme of Equation (16) is used to discretize the convective term of the above governing Equation [27].
φ f , s o u = φ + φ · r
where φ and φ are the cell-centered value and its gradient in the upstream cell and r is the displacement vector from upstream cell centroid to the face centroid.

2.3. Data Reduction

The hydraulic diameter, dh, was defined as Equation (17):
d h = 4 A C w e t
where A is the semicircular cross section area of the channel, and C w e t is the wet circumference of the cross section.
The average heat convection coefficient, h, was defined as Equation (18):
h = Q w T w T b
where Q w is the heat flux of the wall, T w is the average wall temperature, and T b is the fluid bulk temperature.
The channel total pressure drop, Δ P , was defined as Equation (19):
Δ P = P i n P o u t
where P i n and P o u t are area-weighted average pressure at the inlet and outlet of the channel, respectively.

2.4. Grid Independence and Model Validation

Meshes in this study are generated using ICEM CFD 2019, and the size of the first layer adjacent to the wall is less than 2 × 10 6   m to keep the wall y+ < 1. Structured hexahedral cells are used in the whole computational domain, and the mesh is locally densified at the bend of the flow direction, as illustrated in Figure 3.
Mesh independence analysis was performed with five different mesh sizes of 91,656, 152,988, 233,508, 314,028, and 394,548. Percentage changes in physical variables h and Δ P were chosen as the basis for independence judgment. As the comparison result listed in Table 3, the relative error percentage of h changes from 11.01% to 1.17%, and the relative error percentage of Δ P changes from 19.78% to 0.59% with the mesh refinement. Consequently, the mesh size of 314,028 elements is considered sufficient, and 300,000 is used as the baseline for all the rest of the simulations.

2.5. Validation by Experimental Data

The experimental data in [28] were used to verify the accuracy of the numerical method. The experiments were carried out to investigate the thermal performance of sCO2 in a straight channel with semicircular cross section. The hydraulic diameter and total length of the channel are 1.17 mm and 500 mm. The operating pressure, heat flux, and mass flux are 8.1 MPa, −20 kW/m2, and 330 kg/m2-s, respectively. As shown in Figure 4, the CFD simulation results and experimental data maintain a good consistency with the maximum relative error 17.1% over all analyzed Tb range, which covers the pseudocritical point. Thus, the numerical method adopted in this study for the flow and heat transfer performance of sCO2 in the semicircular channel is reliable and comparatively accurate.

3. Results and Discussion

3.1. Effect of Lp on the Channel Flow and Heat Transfer Performance

To study the effect of Lp on thermal and hydraulic performance of zigzag channel, three different Lp (3, 4.5, and 6 mm) are considered for this comparative study. The other geometric factors dh and θ take the valve 1.17 mm and 115°, respectively, and remain unchanged to avoid coupling effects. The channel has an inlet mass flux of G = 200 kg/m2-s, an outlet pressure of Pout = 8 MPa, and a wall heat flux of Qw = ±12 kw/m2-k. The bulk temperature Tb of CO2 varies between 280 K and 360 K, covering the pseudocritical temperature Tm.
As shown in Figure 5, h of the three channels gradually increases and reaches the maximum value as the Tb of the fluid approaches Tm. This is due to the surge of the specific heat and thermal conductivity of CO2 near the pseudocritical point. Δ P of the three types of channels decrease with the increase of Tb. This is mainly because the density of CO2 decreases with the rising of Tb. It can be seen from the comparison results of the three channels that the heat convection coefficient h and pressure drop Δ P both decrease with the rising Lp.
Figure 6 shows the velocity vector along the zigzag channel with different Lp. The flow-field distribution possesses certain periodicity. A large velocity gradient occurs at the channel corner and the maximum velocity in the channel also appears in this area. Boundary layer separation occurs at the corner of the channel and the local velocity increases due to the appearance of vortex. As the velocity direction is different from the wall direction of the next pitch, it can strengthen the mixing of the mainstream and the fluid near the wall, which is conducive to the heat transfer enhancement. As can be seen from the partial enlarged view, the local velocity increases with the reduction of channel Lp, which means that the fluid at the boundary region is mixed with the fluid in the core region more sufficiently. As a result, the reduction of Lp enhances the channel heat transfer, and improves the channel total heat convection coefficient h. The local flow resistance Δ P also rises with the reduction of Lp as the wall separation increases.
Figure 7 represents the local convective heat transfer coefficient in zigzag channel with different Lp. The local heat transfer coefficient increases significantly on the windward side of the corner area. It is because the boundary layer is locally thinner under the direct flushing of the incoming flow. As mentioned above, the local fluid velocity increases with the decrease of channel Lp, which also leads to the local heat transfer coefficient increasing.

3.2. Effect of θ on the Channel Flow and Heat Transfer Performance

In this part of the analysis, four different θ of the channel (100°, 115°, 140°, and straight) are considered for this comparative study with the inlet mass flux G = 200 kg/m2-s, outlet pressure Pout = 8 MPa, and wall heat flux Qw = ±12 kw/m2-k. dh and Lp of the channels remain unchanged with the values 2 mm and 7.75 mm, respectively. The bulk temperature Tb of the CO2 varies between 280 K and 360 K to cover the pseudocritical temperature Tm.
As can be seen from Figure 8, h and Δ P both decrease with the increase of θ, and better thermal performance for all of the zigzag channels is demonstrated, compared with the straight channel. The variation trend of h and Δ P of zigzag channel with Tb is the same as that of the straight channel. This provides an approach for defining the flow and heat transfer correlations in zigzag channels, as there have been several studies on the correlations of Nu of the sCO2 semicircular straight channel [29,30].
Figure 9 shows the comparison of the velocity vector along the channel with different bend angles of the zigzag channel. A sharper bending angle will significantly increase the local fluid velocity at the turning position and aggravate the separation of the boundary layer, which will result in more violent mixing of fluid between the wall area and the core region. It means that the decrease of θ enhances the channel convective heat transfer under the geometric parameters of the current study. Therefore, h increases with the decrease of θ. As θ =180°, the channel becomes a straight channel, and h is smaller than any of the zigzag channels with bending angles.
Figure 10 shows the contour plots of local convective heat transfer coefficient in zigzag channel under different θ. It can be seen that the convective heat transfer coefficient of the wall surface of the zigzag channel is significantly higher than that of the straight channel. The region with the highest local convective heat transfer coefficient appears on the windward side of the turning angle of zigzag channel. It is because this area is washed by the mainstream and has a locally thinner boundary layer.

3.3. Effect of dh on the Channel Flow and Heat Transfer Performance

In this comparative evaluation study, three different dh of 1.17, 2, and 4 mm are used. The other geometric factors Lp and θ are set to 4.5 mm and 115°, respectively and remain unchanged. The channel has an inlet mass flux of G = 200 kg/m2-s, an outlet pressure of Pout = 8 MPa, and a wall heat flux of Qw = ±12 kw/m2-k. The bulk temperature Tb of CO2 varies between 280 K and 360 K, covering the pseudocritical temperature Tm.
With the reduction of channel dh, the distance from the mainstream region to the wall also decreases, which is theoretically conducive to the heat transfer performance, and there are indeed such conclusions in the study on the thermal performance of the straight channel in [24,31]. Nonetheless, for the zigzag channels, the heat convection coefficient h does not show an obvious increasing trend with the decrease of dh. It can be seen from Figure 11 that h and Δ P increase significantly as dh changes from 4 mm to 2 mm, especially in the heating cases. However, when dh changes from 2 mm to 1.17 mm, neither h nor Δ P show significant change. This is different from the conclusion in the semicircular straight channel study.
Through the previous analysis cases, we found that the separation of boundary layer promotes the mixing and diffusion in the fluid and enhances the heat transfer performance of the zigzag channel. However, on the other hand, it will also reduce the effective heat transfer area of the wall, which is disadvantageous to the heat exchange. Figure 12 shows us another possible scenario. As for the dh = 4 mm diameter case, the boundary layer separation area accounts for a large proportion of the total heat exchange area and the weakening effect of the separation of boundary layer on heat transfer becomes dominant. It can also be seen from Figure 13 that the local heat convective coefficient of the dh = 4 mm channel has not been obviously enhanced on the windward side compared to the dh = 1.17 mm and dh = 2 mm channels.
As can be also seen from Figure 12, the boundary layer separation effect is weakened with the decrease of the dh. In the dh = 1.17 mm and dh = 2 mm diameter cases, the boundary layer separation area will not account for as large a proportion of the total wall area as that presented in dh = 4 mm case, which means that the boundary layer separation has a positive effect on the heat transfer performance of the 1.17 mm and 2 mm channels. When Lp >> dh, this positive effect is enhanced with the increase of dh. At the same time, there is an opposite influence whereby the heat transfer will be weakened with the increase of dh due to the increasing distance between mainstream region and the wall. The combination of these two effects makes the convective heat transfer coefficient close for the 1.17 mm and 2 mm channels.
All three geometric parameters produce effects on the flow and heat transfer performance of the zigzag channel and have a certain regularity. When the size of dh is close to Lp, the wall separation caused by channel turning will not strengthen the heat transfer performance. For industrial design, from the point of view of enhancing heat transfer, the Lp should be significantly larger than dh for the zigzag channels.

4. Conclusions

The thermal and hydraulic performance of sCO2 in zigzag channels of different pitch lengths (3 mm, 4.5 mm, 6 mm, 7.75 mm), bending angles (100°, 115°, 140°, straight), and hydraulic diameters (1.17 mm, 2 mm, 4 mm) are studied comparatively with the boundary condition covering the pseudocritical point using the CFD method. In this numerical method, the SST k–ω is adopted as the turbulence model and shows the quantitative prediction of the experiments’ heat transfer data. The following conclusions were obtained:
(1)
The drastic change of CO2 thermophysical parameters has a significant effect on the hydraulic and heat transfer characteristics of the zigzag channel near the pseudocritical point, and its variation tendency with bulk temperature is the same as that of the straight channel.
(2)
The reduction of pitch length enhances the effect of boundary layer separation behind the corner, which can improve the heat transfer performance. As a result, the heat convective coefficient and pressure drop of sCO2 in the zigzag channel increase with the decrease of the pitch length.
(3)
The decrease of channel bend angle can also increase the local velocity at the turning location and enhance the boundary layer separation effect. Therefore, the heat convective coefficient and pressure drop of sCO2 in the zigzag channel both increase with the decrease of channel bend angle.
(4)
The decrease of channel hydraulic diameter is conducive to the heat transfer of the zigzag channel, but it is not as significant as that of the straight channel, because the boundary layer separation effect will be weakened with the decrease of channel hydraulic diameter.

Author Contributions

Conceptualization, Y.T. and Y.Z.; methodology, Y.T.; software, Y.T.; validation, Y.T. and Y.Z.; formal analysis, Y.T.; investigation, Y.T.; resources, Y.Z.; data curation, Y.T.; writing—original draft preparation, Y.T.; writing—review and editing, Y.Z.; visualization, Y.T.; supervision, Y.Z.; project administration, Y.T.; funding acquisition, Y.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Natural Science project of Hunan Province, China, grant number 2020JJ5393 and Education Department of Hunan Province, China, grant number 21B0618.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

ACross section area [m2]
CCircumference [m]
C p Specific heat [J/kg-K]
d Diameter [m]
fFriction factor
GMass flux [kg/m2-s]
hHeat convection coefficient, [W/m2-K]
k Thermal conductivity [W/m-K], turbulent energy [J/kg]
LPPitch length [m]
NuNusselt number
PPressure [Pa]
QHeat flux [W/m2]
ReReynolds number
TTemperature [K]
uFluid velocity [m/s]
yWall distance [m]
ρ Density [kg/m3]
Δ P Pressure drop [Pa]

Greek Symbols

α Heat convection coefficient: [W/m2-K]
ρ Density [kg/m3]
ωSpecific rate of turbulence dissipation [s−1]
μDynamic viscosity [kg/m-s]
yWall distance [m]
θBend angle [°]

Subscripts

bBulk
hHydraulic
inInlet
mPseudocritical point
outOutlet
wWall

References

  1. Chen, M.; Sun, X.; Christensen, R.N.; Shi, S.; Skavdahl, I.; Utgikar, V.; Sabharwall, P. Experimental and Numerical Study of a Printed Circuit Heat Exchanger. Ann. Nucl. Energy 2016, 97, 221–231. [Google Scholar] [CrossRef] [Green Version]
  2. Kim, W.; Baik, Y.J.; Jeon, S.; Jeon, D.; Byon, C. A Mathematical Correlation for Predicting the Thermal Performance of Cross, Parallel, and Counterflow PCHEs. Int. J. Heat Mass Transf. 2017, 106, 1294–1302. [Google Scholar] [CrossRef]
  3. Pua, L.M.; Rumbold, S.O. Industrial Microchannel Devices—Where Are We Today? In International Conference on Microchannels and Minichannels; ASME: New York, NY, USA, 2003; Volume 1, pp. 773–780. [Google Scholar] [CrossRef]
  4. Chen, M. Performance Testing and Modeling of Printed Circuit Heat Exchangers for Advanced Nuclear Reactor Applications. Ph.D. Thesis, University of Michigan, Ann Arbor, MI, USA, 2018. [Google Scholar]
  5. Xu, X.Y.; Wang, Q.W.; Li, L.; Ekkad, S.V.; Ma, T. Thermal-Hydraulic Performance of Different Discontinuous Fins Used in a Printed Circuit Heat Exchanger for Supercritical CO2. Numer. Heat Transf. Part A Appl. 2015, 68, 1067–1086. [Google Scholar] [CrossRef]
  6. Nikitin, K.; Kato, Y.; Ngo, L. Printed Circuit Heat Exchanger Thermal-Hydraulic Performance in Supercritical CO2 Experimental Loop. Int. J. Refrig. 2006, 29, 807–814. [Google Scholar] [CrossRef]
  7. Kruizenga, A.; Anderson, M.; Fatima, R.; Corradini, M.; Towne, A.; Ranjan, D. Heat Transfer of Supercritical Carbon Dioxide in Printed Circuit Heat Exchanger Geometries. In Proceedings of the 14th International Heat Transfer Conference, IHTC 14, Washington, DC, USA, 8–13 August 2010; Volume 4, pp. 653–661. [Google Scholar] [CrossRef] [Green Version]
  8. Saeed, M.; Kim, M.H. Thermal and Hydraulic Performance of SCO2 PCHE with Different Fin Configurations. Appl. Therm. Eng. 2017, 127, 975–985. [Google Scholar] [CrossRef]
  9. Douglas, A.; Olson, D.A. Heat Transfer in Turbulent Supercritical Carbon Dioxide Flowing in a Heated Horizontal Tube; NIST Pubs: Gaithersburg, MD, USA, 1998. [Google Scholar]
  10. Tu, Y.; Zeng, Y. Heat Transfer and Hydraulic Characteristics of Supercritical CO2 in Cooled and Heated Horizontal Semicircular Channels. J. Appl. Fluid Mech. 2021, 14, 1351–1362. [Google Scholar] [CrossRef]
  11. Kim, D.E.; Kim, M.H.; Cha, J.E.; Kim, S.O. Numerical Investigation on Thermal-Hydraulic Performance of New Printed Circuit Heat Exchanger Model. Nucl. Eng. Des. 2008, 238, 3269–3276. [Google Scholar] [CrossRef]
  12. Mohammed, H.A.; Gunnasegaran, P.; Shuaib, N.H. Influence of Channel Shape on the Thermal and Hydraulic Performance of Microchannel Heat Sink. Int. Commun. Heat Mass Transf. 2011, 38, 474–480. [Google Scholar] [CrossRef]
  13. Matsuo, B.K.Y.; Ranjan, D.; Anderson, M. Numerical Study of Compact Heat Exchanger Designs for Generation IV Supercritical Carbon Dioxide Power Conversion Cycles. Nucl. Sci. Eng. 2014, 176, 138–153. [Google Scholar] [CrossRef]
  14. Tsuzuki, N.; Kato, Y.; Ishiduka, T. High Performance Printed Circuit Heat Exchanger. Appl. Therm. Eng. 2007, 27, 1702–1707. [Google Scholar] [CrossRef]
  15. Ngo, T.L.; Kato, Y.; Nikitin, K.; Ishizuka, T. Heat Transfer and Pressure Drop Correlations of Microchannel Heat Exchangers with S-Shaped and Zigzag Fins for Carbon Dioxide Cycles. Exp. Therm. Fluid Sci. 2007, 32, 560–570. [Google Scholar] [CrossRef]
  16. Wen, Z.X.; Lv, Y.G.; Li, Q. Comparative Study on Flow and Heat Transfer Characteristics of Sinusoidal and Zigzag Channel Printed Circuit Heat Exchangers. Sci. China Technol. Sci. 2020, 63, 655–667. [Google Scholar] [CrossRef]
  17. Dai, Z.; Zheng, Z.; Fletcher, D.F.; Haynes, B.S. Experimental Study of Transient Behaviour of Laminar Flow in Zigzag Semi-Circular Microchannels. Exp. Therm. Fluid Science 2015, 68, 644–651. [Google Scholar] [CrossRef]
  18. Chen, M.; Sun, X.; Christensen, R.N. Thermal-Hydraulic Performance of Printed Circuit Heat Exchangers with Zigzag Flow Channels. Int. J. Heat Mass Transf. 2019, 130, 356–367. [Google Scholar] [CrossRef]
  19. Yoon, S.J.; O’Brien, J.; Chen, M.; Sabharwall, P.; Sun, X. Development and Validation of Nusselt Number and Friction Factor Correlations for Laminar Flow in Semi-Circular Zigzag Channel of Printed Circuit Heat Exchanger. Appl. Therm. Eng. 2017, 123, 1327–1344. [Google Scholar] [CrossRef]
  20. Ren, Z.; Zhao, C.R.; Jiang, P.X.; Bo, H.L. Investigation on Local Convection Heat Transfer of Supercritical CO2 during Cooling in Horizontal Semicircular Channels of Printed Circuit Heat Exchanger. Appl. Therm. Eng. 2019, 157, 113697. [Google Scholar] [CrossRef]
  21. Li, H.; Kruizenga, A.; Anderson, M.; Corradini, M.; Luo, Y.; Wang, H.; Li, H. Development of a New Forced Convection Heat Transfer Correlation for CO2 in Both Heating and Cooling Modes at Supercritical Pressures. Int. J. Therm. Sci. 2011, 50, 2430–2442. [Google Scholar] [CrossRef]
  22. Xiang, M.; Guo, J.; Huai, X.; Cui, X. Thermal Analysis of Supercritical Pressure CO2 in Horizontal Tubes under Cooling Condition. J. Supercrit. Fluids 2017, 130, 389–398. [Google Scholar] [CrossRef]
  23. Liu, S.; Huang, Y.; Wang, J. Theoretical and Numerical Investigation on the Fin Effectiveness and the Fin Efficiency of Printed Circuit Heat Exchanger with Straight Channels. Int. J. Therm. Sci. 2018, 132, 558–566. [Google Scholar] [CrossRef]
  24. Tu, Y.; Zeng, Y. Flow and Heat Transfer Characteristics Study of Supercritical CO2 in Horizontal Semicircular Channel for Cooling Process. Case Stud. Therm. Eng. 2020, 21, 100691. [Google Scholar] [CrossRef]
  25. Batchelor, C.K.; Batchelor, G.K. An Introduction to Fluid Dynamics; Cambridge University Press: Cambridge, UK, 2000; ISBN 0521663962. [Google Scholar]
  26. Kamiński, M.; Ossowski, R.L. Navier-Stokes problems with random coefficients by the Weighted Least Squares Technique Stochastic Finite Volume Method. Arch. Civ. Mech. Eng. 2014, 14, 745–756. [Google Scholar] [CrossRef] [Green Version]
  27. Barth, T.J.; Jespersen, D.C. The Design and Application of Upwind Schemes on Unstructured Meshes. In Aiaa Aerospace Sciences Meeting; AIAA: Reston, VA, USA, 1989; p. 0366. [Google Scholar]
  28. Li, H.; Zhang, Y.; Zhang, L.; Yao, M.; Kruizenga, A.; Anderson, M. PDF-Based Modeling on the Turbulent Convection Heat Transfer of Supercritical CO2 in the Printed Circuit Heat Exchangers for the Supercritical CO2 Brayton Cycle. Int. J. Heat Mass Transf. 2016, 98, 204–218. [Google Scholar] [CrossRef]
  29. Kruizenga, A.; Li, H.; Anderson, M.; Corradini, M. Supercritical Carbon Dioxide Heat Transfer in Horizontal Semicircular Channels. J. Heat Transf. 2012, 134, 081802. [Google Scholar] [CrossRef]
  30. Jackson, J.D. Fluid Flow and Convective Heat Transfer to Fluids at Supercritical Pressure. Nucl. Eng. Des. 2013, 264, 24–40. [Google Scholar] [CrossRef]
  31. Dang, C.; Hihara, E. In-Tube Cooling Heat Transfer of Supercritical Carbon Dioxide. Part 2. Comparison of Numerical Calculation with Different Turbulence Models. Int. J. Refrig. 2004, 27, 748–760. [Google Scholar] [CrossRef]
Figure 1. Thermophysical properties of supercritical CO2 at different pressure: (a) specific heat; (b) density; (c) dynamic viscosity; (d) thermal conductivity.
Figure 1. Thermophysical properties of supercritical CO2 at different pressure: (a) specific heat; (b) density; (c) dynamic viscosity; (d) thermal conductivity.
Energies 15 02099 g001
Figure 2. Physical model of zigzag channel.
Figure 2. Physical model of zigzag channel.
Energies 15 02099 g002
Figure 3. Mesh structure of the computational domain.
Figure 3. Mesh structure of the computational domain.
Energies 15 02099 g003
Figure 4. Validation with experimental data [28].
Figure 4. Validation with experimental data [28].
Energies 15 02099 g004
Figure 5. Effect of LP on flow and heat transfer performance: (a) h of cooling case; (b) Δ P of cooling case; (c) h of heating case; (d) Δ P of heating case.
Figure 5. Effect of LP on flow and heat transfer performance: (a) h of cooling case; (b) Δ P of cooling case; (c) h of heating case; (d) Δ P of heating case.
Energies 15 02099 g005
Figure 6. Velocity vector along the zigzag channel with Lp of 3 mm, 4.5 mm, and 6 mm.
Figure 6. Velocity vector along the zigzag channel with Lp of 3 mm, 4.5 mm, and 6 mm.
Energies 15 02099 g006
Figure 7. Local heat transfer coefficient of the zigzag channel with Lp of 3 mm, 4.5 mm, and 6 mm.
Figure 7. Local heat transfer coefficient of the zigzag channel with Lp of 3 mm, 4.5 mm, and 6 mm.
Energies 15 02099 g007
Figure 8. Effect of θ on flow and heat transfer performance: (a) h of cooling case; (b) Δ P of cooling case; (c) h of heating case; (d) Δ P of heating case.
Figure 8. Effect of θ on flow and heat transfer performance: (a) h of cooling case; (b) Δ P of cooling case; (c) h of heating case; (d) Δ P of heating case.
Energies 15 02099 g008
Figure 9. Velocity vector along the zigzag channel with θ of 100°, 115°, 140°, and straight.
Figure 9. Velocity vector along the zigzag channel with θ of 100°, 115°, 140°, and straight.
Energies 15 02099 g009
Figure 10. Local heat transfer coefficient of the zigzag channel with θ of 100°, 115°, 140°, and straight.
Figure 10. Local heat transfer coefficient of the zigzag channel with θ of 100°, 115°, 140°, and straight.
Energies 15 02099 g010
Figure 11. Effect of dh on flow and heat transfer performance: (a) h of cooling case; (b) Δ P of cooling case; (c) h of heating case; (d) Δ P of heating case.
Figure 11. Effect of dh on flow and heat transfer performance: (a) h of cooling case; (b) Δ P of cooling case; (c) h of heating case; (d) Δ P of heating case.
Energies 15 02099 g011
Figure 12. Velocity vector along the zigzag channel with dh of 1.17 mm, 2 mm, and 4 mm.
Figure 12. Velocity vector along the zigzag channel with dh of 1.17 mm, 2 mm, and 4 mm.
Energies 15 02099 g012
Figure 13. Local heat transfer coefficient of the zigzag channel with dh of 1.17 mm, 2 mm, and 4 mm.
Figure 13. Local heat transfer coefficient of the zigzag channel with dh of 1.17 mm, 2 mm, and 4 mm.
Energies 15 02099 g013
Table 1. Geometric parameters.
Table 1. Geometric parameters.
Geometric ParametersValues
Channel hydraulic diameters, dh, mm1.17 (experimental validation), 2, 4
Bending angle, θ, °100, 115, 140, Straight
Channel pitch length, Lp, mm3, 4.5, 6, 7.75
Channel total length, L, mm500
Table 2. Boundary conditions in detail.
Table 2. Boundary conditions in detail.
InletOutletWall
Temperature (K)Mass flux (kg/m2-s)Pressure (MPa)Constant heat flux (kW/m2)
280~3502007.5, 8, 9±12
Table 3. Mesh independence analysis.
Table 3. Mesh independence analysis.
No.Number of Elementsh (W/m2-K)Error (%) Δ P ( Pa ) Error (%)
191,6561655.29811.01%629.719.78%
2152,9881644.92611.56%639.921.72%
3233,5081737.7786.57%605.915.26%
4314,0281838.2271.17%528.80.59%
5394,5481860.0270.00%525.70.00%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tu, Y.; Zeng, Y. Numerical Study on Flow and Heat Transfer Characteristics of Supercritical CO2 in Zigzag Microchannels. Energies 2022, 15, 2099. https://doi.org/10.3390/en15062099

AMA Style

Tu Y, Zeng Y. Numerical Study on Flow and Heat Transfer Characteristics of Supercritical CO2 in Zigzag Microchannels. Energies. 2022; 15(6):2099. https://doi.org/10.3390/en15062099

Chicago/Turabian Style

Tu, Yi, and Yu Zeng. 2022. "Numerical Study on Flow and Heat Transfer Characteristics of Supercritical CO2 in Zigzag Microchannels" Energies 15, no. 6: 2099. https://doi.org/10.3390/en15062099

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