Next Article in Journal
A Design of a Hybrid Non-Linear Control Algorithm
Previous Article in Journal
Wind Turbine Loads Induced by Terrain and Wakes: An Experimental Study through Vibration Analysis and Computational Fluid Dynamics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Investigation on the Flow and Heat Transfer Characteristics of Supercritical Liquefied Natural Gas in an Airfoil Fin Printed Circuit Heat Exchanger

School of Energy and Power, Jiangsu University of Science and Technology, Zhenjiang 210000, China
*
Author to whom correspondence should be addressed.
Energies 2017, 10(11), 1828; https://doi.org/10.3390/en10111828
Submission received: 25 July 2017 / Revised: 23 October 2017 / Accepted: 1 November 2017 / Published: 10 November 2017
(This article belongs to the Special Issue Liquefied Natural Gas)

Abstract

:
As a new kind of highly compact and efficient micro-channel heat exchanger, the printed circuit heat exchanger (PCHE) is a promising candidate satisfying the heat exchange requirements of liquefied natural gas (LNG) vaporization at low and high pressure. The effects of airfoil fin arrangement on heat transfer and flow resistance were numerically investigated using supercritical liquefied natural gas (LNG) as working fluid. The thermal properties of supercritical LNG were tested by utilizing the REFPROF software database. Numerical simulations were performed using FLUENT. The inlet temperature of supercritical LNG was 121 K, and its pressure was 10.5 MPa. The reference mass flow rate of LNG was set as 1.22 g/s for the vertical pitch Lv = 1.67 mm and the staggered pitch Ls = 0 mm, with the Reynolds number of about 3750. The SST k-ω model was selected and verified by comparing with the experimental data using supercritical liquid nitrogen as cold fluid. The airfoil fin PCHE had better thermal-hydraulic performance than that of the straight channel PCHE. Moreover, the airfoil fins with staggered arrangement displayed better thermal performance than that of the fins with parallel arrangement. The thermal-hydraulic performance of airfoil fin PCHE was improved with increasing Ls and Lv. Moreover, Lv affected the Nusselt number and pressure drop of airfoil fin PCHE more obviously. In conclusion, a sparser staggered arrangement of fins showed a better thermal-hydraulic performance in airfoil fin PCHE.

1. Introduction

Owing to its high calorific value and low carbon dioxide emissions, natural gas (NG) has become the best choice for replacing traditional resources such as coal and petroleum [1,2]. Generally; NG is cooled to form liquefied natural gas (LNG), allowing it to be stored and transported over long distances. LNG is then heated and regasified into NG before being transferred into pipelines to meet the demand of users [3,4]. Efficient and reliable vaporization devices have been highlighted as critical equipment in the LNG gasification system. Four main kinds of LNG vaporizers are are commercially available: intermediate fluid vaporizers; open rack vaporizers (ORVs); super ORVs and submerged combustion vaporizers [5,6,7].
However, traditional LNG vaporizers cannot meet the requirements of miniaturization and high compactness in the construction of LNG gasification systems. Therefore, it is urgent to find an efficient and compact heat exchanger to replace these traditional vaporizers. Because of safety, high compactness and efficiency, printed circuit heat exchangers (PCHEs) have drawn wide attention in recent years [8,9,10,11]. A PCHE is a micro-channel heat exchanger manufactured by photochemical etching and diffusion bonding [12,13,14,15]. As a result, PCHEs show better heat transfer performance than those of other types of heat exchanger due to their larger heat surface area caused by the presence of many micro-channels. Nevertheless, PCHEs suffer from enhanced pressure drop. Therefore, researchers have endeavored to improve the thermal-hydraulic characteristics of PCHEs [16,17,18,19,20,21,22,23].
Four types of PCHE flow channels, i.e., straight, zigzag, S-shape, and airfoil fin have been developed so far. Mylavarapu et al. [24] experimentally tested PCHEs with straight, circular and semicircular channels at 2.7 MPa with various Reynolds numbers. The transition from laminar flow to flow regime occurred much earlier in the semicircular channel than within the circular pipe.
Ma et al. [25] studied the pressure drop mechanism and the local heat transfer of PCHEs with zigzag channels and the effect of the inclination angle of channels on the thermal-hydraulic performance of the PCHEs under the operating conditions of a very high temperature reactor. Figley et al. [26] studied the heat transfer and pressure drop characteristics of PCHEs with zigzag channels by the aid of three-dimensional numerical simulations. Aneesh et al. [27] studied the thermal-hydraulic characteristics and properties of 3D straight channel-based PCHEs. Khan et al. [28] investigated the thermal-hydraulic characteristics and performance of PCHEs for various angles of bend (θ = 0°, 5°, 10°, and 15°) and different Reynolds numbers (Re = 350, 700, 1400 and 2100). As Re and θ increased, the thermal hydraulic performance was boosted.
PCHEs with straight and Z-shaped channels were analyzed initially. However, many researchers have found that PCHEs with distributed fins, such as S-shaped fins and airfoil fins, have better thermal-hydraulic characteristics. Ngo et al. [29] studied a new PCHE with an S-shaped fin structure applied to a hot water supplier which uses supercritical CO2 as heat source. The new PCHE had about 3.3 times less volume, but it displayed a 37% lower pressure drop on the CO2 side and by ten times on the H2O side than an existing 1.5 MW/m3 hot water supplier.
Kim et al. [30] proposed an airfoil fin, making the pressure drop of the airfoil fin one-twentieth of that of zigzag channel, but their total heat transfer rates per unit volume were basically the same, to enhance the thermal hydraulic performance of PCHE. Although PCHEs have better thermal hydraulic performance than traditional heat exchangers due to the smaller hydraulic diameter of their flow channels, this also raises the pressure drop. In addition to optimizing the channel shape of PCHEs, many researchers have focused on the fluid in the flow channels [31]. Compared to traditional fluids, supercritical fluids have many advantageous characteristics as the working fluid in PCHEs, such as high density, low viscosity and high thermal conductivity [32,33,34].
Supercritical carbon dioxide and water have been applied in PCHEs. Kruizenga et al. [35] performed an experimental study to estimate the pressure drop and heat transfer of PCHEs in which the supercritical CO2 was selected as the working fluid. The heat transfer significantly increased when the pseudocritical temperature and critical pressure were approached.
Kim et al. [36] performed numerical investigations to study the performance of airfoil fin PCHEs in which supercritical carbon dioxide was used as the working fluid. They examined the optimal arrangement of airfoil fins with an objective function, and the arrangement had a staggered spacing of 1 mm.
Despite the extensive studies on the thermal hydraulic performance of PCHEs using supercritical carbon dioxide and water as working fluids, the thermal-hydraulic performance of PCHEs used in an LNG vaporization system at low temperature and high pressure has rarely been reported hitherto. In this study, the numerical simulation method was used to analyze the thermal-hydraulic performance of a PCHE using supercritical LNG as the work fluid. The heat transfer and flow characteristics of supercritical LNG in a single channel of straight channel PCHE and airfoil fin PCHE were compared at different mass fluxes. To obtain the correlations between performance factors and configuration factors, pressure drop and heat transfer were replaced with dimensionless representations (Nusselt number and Euler number). The effects of airfoil fin arrangements on heat transfer and pressure drop characteristics were analyzed. Finally, an optimal design for the fin arrangement in an airfoil fin PCHE was suggested. The results in favour of the optimum thermal design and operation of high-performance airfoil fin PCHE.

2. Computational Fluid Dynamic Analysis for Thermal-Hydraulic Performance

2.1. Thermal–Physical Properties of Supercritical LNG

In this study, the LNG pressure was 10.5 MPa, exceeding its critical pressure (P = 4.59 MPa). As a result, LNG in the PCHE channel was in the supercritical state, whereby an obvious liquid phase and gas phase do not exist. Like other supercritical fluids, supercritical LNG also has various favorable characteristics like high density, high thermal conductivity and low viscosity. The properties including density, thermal conductivity, specific heat and viscosity were calculated by REFPROP 9.0 software (Figure 1). Meanwhile, the accuracy of the property data in the entire regime computation was validated (Figure 2). The errors were within ±4%, indicating the viability of this computation method. The thermal properties of supercritical LNG were calculated with the FLUENT software by defining the piecewise-linear functions of temperature (Table 1).

2.2. Physical Models and Definition of Airfoil Fin Arrangement Parameters

Generally, an airfoil fin PCHE consists of several micro-channels, with several airfoil fins in one channel. Therefore, it is extremely difficult to simulate the entire heat exchanger geometry because a powerful computer and long computation time are needed. However, as shown in Figure 3, the flow channel formed by airfoil fins is periodic in both the longitudinal direction and the transverse direction, except for the plate border in a PCHE. Therefore, considering the computing resource, forty periodic fin structures along the flow direction from the core area of heat exchanger with the length of 260 mm were selected firstly. The vertical pitch Lv and staggered pitch Ls of airfoil fins were 1.67 mm and 0 mm respectively, aiming to compare the heat transfer and flow characteristics of supercritical LNG between airfoil fin and straight channel with the same channel length and hydraulic diameter. The numerical models of straight channel and airfoil fin are shown in Figure 4 and Figure 5, respectively.
Then, to study the effects of airfoil fin arrangements on the flow and heat transfer characteristics of supercritical LNG, six periodic fin structures along the flow direction and three along the transverse direction were selected as the simulation domain. The airfoil fin PCHE has three important geometric parameters (Figure 3). The separation distance of staggered arrangement is presented by Ls. There is a periodicity between an un-staggered arrangement and a fully staggered arrangement. Lh indicates the separation distance between one airfoil head and adjacent airfoil head in a row. The separation distance between one row and adjacent row in the vertical direction is indicated by Lv. Herein, Lh was 6 mm, Ls varied from 0 to 4 mm and Lv changed from 1.3 mm to 3 mm. According to the configuration, the width (W) of the heat transfer region of the entire domain varied from 3.9 mm to 9 mm. The vertical pitch Lv = 1.67 mm and staggered pitch Ls = 0 mm were selected as the baseline model.
A schematic diagram of the simulation domain (Figure 6) showed that Lv = 1.67 mm and Ls = 0 mm. The working fluid was supercritical LNG, and the material of substrate plates and fins was steel. The mass flux inlet was applied for the inlet boundary condition of the simulation domain, and the inlet temperature and reference mass flux were set at 121 K and 325 kg/m2∙s, respectively. The outlet boundary condition was set as pressure outlet. As shown in Figure 7, a periodic boundary condition in the left/right positions, as well as a constant heat flux applied to the top and bottom positions is 58,713.75 W/m2.

2.3. Numerical Method and Grid Independence

ANSYS Fluent 14.5 was used for our numerical evaluation of the airfoil fin PCHE. The set mass flow rate corresponded to turbulent flow regimes in the PCHE cold channels. It is important to select a proper turbulence model for numerical calculations. Kim and No [10] reported that the shear stress transport k-ω model (SST k-ω) [11] predicted a fine grid near the wall in supercritical calculation extremely well. Therefore, the SST k-ω model was used in this study. A detailed description of the SST k-ω model can be found in [37]. The semi-implicit method pressure linked equation algorithm was used to resolve the coupling of velocity and pressure. Before solution convergence, the residual for every variable was required to be less than 10−6. Meanwhile, the second order upwind, which has a smaller truncation error than that of the first order upwind, was used in the momentum equation and energy equation to ensure the accuracy of simulation.
The structured computation mesh was generated by GAMBIT. The number of meshes was different for all cases on account of various dimensions of the simulation domain, so the model with staggered number Ls = 0 mm and vertical number Lv = 1.67 mm was selected as the baseline model. The mesh dependence test dominated the density of meshes (Figure 8). The influence of grid density on the accuracy of calculated results was studied by comparing six sets of grid numbers: 1302567, 1762753, 2415689, 3269854, 3594425 and 4196856 cells. By comparing the outlet temperature and Nusselt number, the optimal grid of 3594422 was selected considering the accuracy and computational efficiency. As shown in Figure 9, the grids are encrypted near the wall surfaces of fins and substrate plates to ensure that y+ is lower than 1. Six boundary layers were established near the top and bottom walls and fin surfaces because of the strict requirements for boundary grid in the supercritical flow and heat transfer, and the thickness of the first boundary layer was 0.01 mm.

2.4. Model Validation

In the present study, the heat transfer and flow characteristics of supercritical LNG in PCHE were investigated numerically. The results of CFD analysis should be compared with the experimental data for validating analysis methodology. However, due to the flammable and explosive properties of LNG when the gas leaks or ventilation is insufficient at high pressure, supercritical nitrogen was selected to substitute LNG as the cold fluid. A crossflow PCHE with airfoil fin channels on the cold side and straight channels on the hot side was manufactured using stainless steel 316L (Figure 10). The airfoil fin with the chord length of 4 mm and the maximum thickness of 0.8 mm was selected. A experimental system was established to study the thermal-hydraulic performance of airfoil fin PCHE using supercritical nitrogen as the cold fluid and R22 as the hot fluid (Figure 11). The inlet temperature of nitrogen was 102 K when its pressure was varied from 5.5 MPa to 7.5 MPa. The mass flux of nitrogen varied from 233 to 421 kg/m2·s, corresponding to turbulent flow regimes on the cold side of airfoil fin PCHE. In short, this PCHE had a good heat transfer ability at high pressure and low temperature. The overall heat transfer coefficient of this PCHE ranged from 850 to 2600 W/m2K, and the heat transfer efficiency was up to approximately 98%.
Figure 10 displays the schematic diagram for the internal core structure of airfoil fin PCHE. It is rather complex to simulate the heat transfer processes of the total tested airfoil fin PCHE, and this study focused on the heat transfer and flow performance of cold fluid. Therefore, a single airfoil fin channel of cold side with full length (400 mm) which used supercritical nitrogen as the working fluid was selected to simulate and to compare with the experimental data (Figure 3). The numerical model of this section was the same as that in Figure 5, except for the length. The inlet temperature of nitrogen was 102 K when its pressure was varied from 5.5 MPa to 7.5 MPa. The mass flux of nitrogen was 421 kg/m2·s, corresponding to turbulent flow regimes on the cold side of airfoil fin PCHE. The differences in the pressure drop and outlet temperature of the cold side between the numerical solutions and the experimental data were compared using Equation (1):
E r r o r = C F D E x p e r i m e n t E x p e r i m e n t × 100 %
The pressure drop differences between the experimental and simulation results were analyzed when the mass flux was 421 kg/m2·s (Table 2). The numerical results differed from the experimental data by 6.9% on average and by 11.62% at maximum. The deviation may be attributed to the inlet and outlet header pressure drop, as well as the uncertainty of pressure transmitters. Besides, the difference between numerical and experimental outlet temperatures on the cold side had an average error of 0.735% and the maximum difference of 1.127%. Therefore, the numerical data were in fairly good agreement with the experimental data, and the numerical model and method used herein were reliable.
However, Figure 1 shows that the physical properties of supercritical fluid change substantially with rising temperature during heat transfer. The local heat transfer and flow characteristics of supercritical fluid cannot be obtained in the experimental apparatus. Therefore, the local parameters of the airfoil fin PCHE were obtained through numerical simulation.

3. Objective Function Parameters

As the characteristic length of the channel, hydraulic diameter Dh is an important value in dimensionless analysis. It is defined as four times the cross-sectional area over a perimeter in the straight channel. Owing to the continual changes of cross-sectional area and perimeter, the hydraulic diameter cannot be the same as that of straight channel, as shown in Figure 12. Nevertheless, the airfoil fin placement in Figure 3 is periodic. Therefore, the hydraulic diameter can be defined by the equations below [37].
V = ( L W S a ) t
S = 2 ( P a t 2 ) + 2 ( L L c ) t + 2 ( W L S a )
D h = 4 V / S
where Sa is the top area of airfoil fin, and Pa is the perimeter of airfoil fin. V and S indicate the volume and side surface area of flow channel, respectively.
The convective heat transfer coefficient can be obtained by the following equation:
h = q w T w a l l T b = q w T w a l l ( T o u t + T i n ) / 2
where q w is the area-averaged wall heat flux, Twall is the area-average wall temperature, and Tin, Tout and Tb are the inlet, outlet and bulk temperatures of LNG obtained from the Fluent data respectively.
The Nusselt number is defined below:
N u = h D h λ
where Dh is the channel hydraulic diameter, and λ is the thermal conductivity of LNG.
The Reynolds number can be calculated by Equation (7):
R e = υ ρ D h μ
where Dh is the channel hydraulic diameter, ρ is the density of LNG and μ is the dynamic viscosity of LNG.
The friction and acceleration effects caused by the density difference between inlet and outlet can result in the pressure drop:
Δ P = Δ P a c c + Δ P f r i c
Δ P a c c = G 2 ( 1 ρ o u t 1 ρ i n )
Δ P f r i c = 2 f L u ρ υ 2 D h
where Δ P is the total pressure drop, Δ P a c c is the acceleration pressure drop and Δ P f r i c is the friction pressure drop.
The Fanning friction factor f is defined as below:
f = τ w 1 2 ρ u m 2
where τ w is the wall shear stress.
In terms of pressure drop, the Euler number was selected as a representative pressure loss coefficient in this study. It reflects the relationship between pressure drop and dynamic velocity head, as well as the relative momentum loss rate:
E u = Δ P ρ u 2 / 2

4. Results and Discussion

4.1. Comparison of Straight Channel PCHE and Airfoil Fin PCHE

In the present study, the heat transfer and flow characteristics of supercritical LNG in a straight channel PCHE and an airfoil fin PCHE with the same hydraulic diameter (Dh = 0.917 mm) were numerically investigated. Figure 13 shows the velocity contours of supercritical LNG in airfoil fin and straight channel PCHEs when the mass flux is 325 kg/m2·s. When supercritical LNG was gradually heated, the bulk velocity significantly increased because of reduced density. However, the velocity increased rapidly in airfoil fin channel owing to continuous expansion and contraction in the sectional area of flow channel. The maximum velocity in the narrowest flow channel was nearly 3 times that in the inlet. Therefore, airfoil fins evidently disturbed supercritical LNG, which increased heat transfer and flow resistance simultaneously.
To evaluate the effect of mass flux, six different mass fluxes were tested for straight channel and airfoil fin PCHEs. Figure 14 display the changes of heat transfer coefficient and Nusselt number at various mass fluxes respectively. Since the turbulence intensity of flow significantly increased with rising mass flux, the heat transfer coefficient and Nusselt number rapidly increased in both PCHEs. Also, the Nusselt number of airfoil fin PCHE exceeded that of straight channel PCHE. As we all know, the heat transfer and flow of a heat exchanger were affected by fins predominantly in two ways: giving rise to disturbance and enlarging heat transfer area. Therefore, the airfoil fin PCHE showed better thermal performance than that of the straight channel PCHE at the same mass flux and hydraulic diameter. For example, when mass flux G = 425 kg/m2∙s and G = 725 kg/m2∙s, Nu values of airfoil fin PCHE were 1.48 and 1.36 times those of the straight channel PCHE, respectively.
The effects of mass flux on the Euler number and Fanning friction factor of the straight channel PCHE and the airfoil fin PCHE were also assessed (Figure 15). The Euler number of the straight channel PCHE surpassed that of the airfoil fin PCHE. For example, when mass flux G = 425 kg/m2∙s, the Euler number of the airfoil fin PCHE was 89.7% of that of the straight channel PCHE. Generally, pressure drop increases with rising mass flux. However, Figure 16 shows that the Euler number calculated by Equation (12) decreases as the mass flux increases. According to this equation, the Euler number is proportional to pressure drop Δ P but inversely proportional to v2. Obviously, the velocity of supercritical LNG rose with increasing mass flux, although it also raised the pressure drop. Nevertheless, v2 exerted a stronger effect on Eu than on pressure drop, so the increase of mass flux led to decrease of the Euler number.
The heat transfer quantity and pressure drop, as the main indices of heat exchanger performance, were described as the Nusselt number and the Euler number herein, both helping to optimize the arrangement of the airfoil fins. Generally, a specific objective function is used to verify an optimal design. The ratio of Nusselt number to Euler number, i.e., Nu/Eu, was employed as the objective function to evaluate the performance of the heat exchanger. Figure 16 describes Nu/Eu as a function of mass flux. The minimum and maximum differences of Nu/Eu between the straight channel PCHE and the airfoil fin PCHE were 46.2% and 51.07%, respectively. Evidently, the airfoil fin PCHE had a better thermal-hydraulic performance, and the difference of Nu/Eu between the airfoil fin and straight channel PCHEs increased with rising mass flux. Hence, fin arrangement is recommended for the airfoil fin PCHE.

4.2. Effect of Fin Arrangement: Staggered Pitch (Ls)

Fins affect flow and heat transfer primarily through their shapes and arrangements in a PCHE. In this section, airfoil fins were arranged in parallel (Ls = 0 mm) and staggered (Ls = 1~4 mm) in the transverse direction of the PCHE.
In an airfoil fin PCHE, the airfoil fins with wide head and narrow tail favor the formation of a much smoother flow channel when arranged in a staggered manner, and the vortex separation could also be prevented by the streamlined fins, which may result in a smaller flow resistance in staggered arrangement than that in parallel arrangement. Figure 17 depicts the velocity contour of supercritical LNG in airfoil fin PCHE arranged in parallel and staggered manners when the vertical pitch Lv = 1.67 mm. In the start of heating, the velocities of fluid were quite uniform and small in both staggered and parallel arrangements. However, owing to continuous expansion and contraction in the sectional area of flow channel, the non-uniformity of velocity became obvious in parallel arrangement as the velocity increased in the flow direction. Therefore, a staggered arrangement of airfoil fins benefited the formation of smooth flow channel and the improvement of flow field uniformity.
Figure 18a shows the Nusselt numbers of an airfoil fin PCHE in parallel and staggered arrangements. Clearly, the airfoil fins with staggered arrangement showed better thermal performance than that of the airfoil fins with parallel arrangement. However, at the same vertical pitch Lv, the Nusselt numbers of parallel and staggered arrangements were similar. The Nusselt number at Ls = 0 mm differed from that at Ls = 4 mm by 6.9% when the vertical number Lv = 3 mm.
In addition, the flow resistance increased as the staggered pitch Ls decreased. Figure 18b,c present the pressure drop and Euler number with the staggered pitch (Ls) at different Lv in an airfoil fin PCHE. In Figure 18c, the Euler number in staggered arrangement (Ls = 1 mm) is smaller (a maximum decrease of 9.86%) than that in parallel arrangement (Ls = 0 mm), with the difference enlarging as the staggered pitch (Ls) increases at the same vertical pitch Lv.
Figure 18d presents the dependence of Nu/Eu on Ls. Nu/Eu increased with rising staggered number Ls. Ls = 0 mm had a lower Nu/Eu (about 27%) than that at Ls = 4 mm, suggesting that Ls = 4 mm had better heat transfer and pressure drop. Probably, pressure drop was more susceptible to Ls than heat transfer. Collectively, staggered arrangement was superior to parallel arrangement in airfoil fin PCHE, manifested as reduced flow resistance and improved total thermal-hydraulic performance.

4.3. Effect of Fin Arrangement: Vertical Pitch (Lv)

When Lh (6 mm) is kept constant, Lv directly determines the density of fins and the width of flow channel. Therefore, the effect of vertical separation distance (Lv) on the thermal-hydraulic performance of an airfoil fin PCHE was assessed.
The velocity contours of supercritical LNG in airfoil fin PCHE arranged at different Lv when the staggered pitch Ls = 1.67 mm are presented in Figure 19. A smaller vertical pitch Lv resulted in a higher velocity, thereby augmenting the flow resistance. In fact, a smaller Lv led to a narrower sectional area of flow channel, so that the flow velocity increased with decreasing Lv at the same mass flux. The maximum velocities of LNG in the narrowest sectional of flow channel were 1.35 m/s and 1.79 m/s in the airfoil fins with Ls = 3 mm and Ls = 1.3 mm, respectively. Therefore, the turbulence intensity of fluid and the flow resistance were both enhanced locally.
Figure 20 shows the variations of heat transfer coefficient h and Nusselt number (Nu) with increasing fin distance (Lv) at different Ls in an airfoil fin PCHE. The convective heat transfer coefficient h decreased with rising Lv at the same Ls, which may attributed to the decreased flow velocity. However, h and Nu changed oppositely with increasing Lv, mainly because the hydraulic diameter increased significantly faster than h decreased with rising Lv. For example, the hydraulic diameter at Lv = 3 mm was nearly 2.7 times that at Lv = 1.3 mm, and h at Lv = 1.3 mm was 1.2 times that at Lv = 1.3 mm when Ls = 4 mm. Thus, the Nusselt number increased with rising Lv of airfoil fins.
Figure 21a shows the dependence of the Euler number on Lv. The total pressure drop decreased as Lv increased, so the Euler number dropped more apparently with increasing Lv at the same Ls. The Euler number at Lv = 3 mm was only 56% of that at Lv = 1.3 mm. Obviously, at the same Ls, reducing Lv slightly facilitated heat transfer, though it also considerably elevated pressure drop.
As described in Section 4.1, a specific objective function Nu/Eu was used to indicate the performance of the airfoil fin PCHE. As illustrated in Figure 21b, Nu/Eu noticeably rises as Lv increases. Nu/Eu at Lv = 3 mm is nearly 3 times and 1.7 times those at Lv = 1.3 mm and Lv = 2 mm respectively, indicating that a dense fin arrangement was conducive to increasing the heat transfer rate. Meanwhile, it was inevitably more difficult to overcome flow resistance. Accordingly, fins should be sparsely arranged in an airfoil fin PCHE.

5. Conclusions

We here in aimed to clarify the thermal-hydraulic characteristics and optimal structure of a PCHE in which supercritical LNG was used as the working fluid. The effects of channel shape and fin arrangements on the flow resistance and heat transfer in the airfoil fin PCHE were numerically investigated as follows:
(1)
The numerical model and methods were validated with experimental data. Supercritical liquid nitrogen was used as a cold fluid for simulation and experiment. The SST model followed by the enhanced wall treatment method well predicted the outlet temperature and pressure drop of a single airfoil fin in the PCHE. The error between the numerical and experimental data was within 14%, indicating the heat transfer and flow characteristics of supercritical LNG in airfoil fin PCHE could be reliably simulated by the model and method.
(2)
As a new type of discontinuous fins, airfoil fins can boost the thermal-hydraulic performance compared with that of a straight channel PCHE using supercritical LNG as the working fluid. The minimum and maximum differences of Nu/Eu between straight channel and airfoil fin PCHEs were 46.2% and 51.07%, respectively. The convective heat transfer coefficient and pressure drop increased in both PCHEs with rising mass flux.
(3)
A staggered fin arrangement was more beneficial to the thermal-hydraulic performance of the airfoil fin PCHE than a parallel fin arrangement using supercritical LNG as the working fluid. At the same Lv and Lh, airfoil fins arranged at Ls = 4 mm displayed better thermal-hydraulic performance than those of the fins at other Ls.
(4)
The velocity of supercritical LNG in the airfoil fin channels increased along the channel length and then plummeted with increasing Lv. The effect of vertical number Lv on the thermal-hydraulic performance of airfoil fin PCHE was more evident than that of staggered pitch Ls. Based on a comprehensive analysis of heat transfer coefficient and pressure drop, a sparser staggered arrangement of fins can enhance the thermal-hydraulic performance of an airfoil fin PCHE.

Acknowledgments

The authors gratefully acknowledge that this work was Jiangsu marine and fishery science and technology innovation and extension project (HY2017-8) and Zhenjiang funds for the key research and development project (GY2016002-1).

Author Contributions

All authors contributed to the paper. Zhongchao Zhao was the director of the research project, put forward the study ideas and wrote this paper; The numerical simulation and data analysis was completed by Kai Zhao and Dandan jia; Pengpeng Jiang and Rendong Shen discussed experimental ideas and completed a part of the data analysis; the laboratory tests were completed by all authors.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

fFanning factor
vVelocity (m/s)
ReReynolds number
hConvective heat transfer coefficient (W/m2∙K)
NuNusselt number
CpSpecific heat (kJ/kg∙K)
DhHydraulic diameter (mm)
q m Mass flow rate (kg/s)
uVelocity (m/s)
q’’Heat flux (W/m2)
GMass flux (kg/m2∙s)
LvThe pitch between one airfoil head and adjacent row in the vertical direction (mm)
LsThe pitch of the staggered arrangement (mm)
LhThe pitch between one airfoil head and adjacent airfoil head in a row (mm)
Δ P Pressure drop (Pa)
Δ p f r i c Pressure drop due to friction (Pa)
Δ p a c c Pressure drop due to acceleration (Pa)
ρ i n Density at the inlet of the channel (kg/m3)
ρ o u t Density at the outlet of the channel (kg/m3)
τ w Shear stress at the wall (Pa)
Greek symbols
μ Viscosity [Pa·s]
ρ Density [kg/m3]
λ Thermal conductivity [W/m2∙K]
Subscript
wWall
bBulk mean
accAcceleration
fricFriction
vVertical
sStaggered
hHorizontal
inInlet
outOutlet

References

  1. Yousefi, A.; Birouk, M. Investigation of natural gas energy fraction and injection timing on the performance and emissions of a dual-fuel engine with pre-combustion chamber under low engine load. Appl. Energy 2017, 189, 492–505. [Google Scholar] [CrossRef]
  2. Pham, T.N.; Long, V.N.D.; Lee, S.; Lee, M. Enhancement of single mixed refrigerant natural gas liquefaction process through process knowledge inspired optimization and modification. Appl. Therm. Eng. 2017, 110, 1230–1239. [Google Scholar] [CrossRef]
  3. Pu, L.; Qu, Z.; Bai, Y.; Qi, D.; Song, K.; Yi, P. Thermal performance analysis of intermediate fluid vaporizer for liquefied natural gas. Appl. Therm. Eng. 2014, 65, 564–574. [Google Scholar] [CrossRef]
  4. Wang, T.; Lin, B. China’s natural gas consumption peak and factors analysis: A regional perspective. J. Clean. Prod. 2017, 142, 548–564. [Google Scholar] [CrossRef]
  5. Pan, J.; Li, R.; Lv, T.; Wu, G.; Deng, Z. Thermal performance calculation and analysis of heat transfer tube in super open rack vaporizer. Appl. Therm. Eng. 2016, 93, 27–35. [Google Scholar] [CrossRef]
  6. Liu, S.; Jiao, W.; Wang, H. Three-dimensional numerical analysis of the coupled heat transfer performance of LNG ambient air vaporizer. Renew. Energy 2016, 87, 1105–1112. [Google Scholar] [CrossRef]
  7. Han, C.-L.; Ren, J.J.; Wang, Y.Q.; Dong, W.-P.; Bi, M.S. Experimental investigation on fluid flow and heat transfer characteristics of a submerged combustion vaporizer. Appl. Therm. Eng. 2017, 113, 529–536. [Google Scholar] [CrossRef]
  8. Yoon, S.-J.; Sabharwall, P.; Kim, E.-S. Numerical study on crossflow printed circuit heat exchanger for advanced small modular reactors. Int. J. Heat Mass Transf. 2014, 70, 250–263. [Google Scholar] [CrossRef]
  9. Tsuzuki, N.; Kato, Y.; Ishiduka, T. High Performance Printed Circuit Heat Exchanger. Appl. Therm. Eng. 2007, 27, 1702–1707. [Google Scholar] [CrossRef]
  10. Kim, I.H.; No, H.C. Thermal hydraulic performance analysis of a printed circuit heat exchanger using a helium–water test loop and numerical simulations. Appl. Therm. Eng. 2011, 31, 4064–4073. [Google Scholar] [CrossRef]
  11. Kim, I.H.; No, H.C.; Lee, J.I.; Jeon, B.G. Thermal hydraulic performance analysis of the printed circuit heat exchanger using a helium test facility and CFD simulations. Nucl. Eng. Des. 2009, 239, 2399–2408. [Google Scholar] [CrossRef]
  12. Kim, I.H.; No, H.C. Physical model development and optimal design of PCHE for intermediate heat exchangers in HTGRs. Nucl. Eng. Des. 2012, 243, 243–250. [Google Scholar] [CrossRef]
  13. Mylavarapu, S.K.; Sun, X.D.; Christensen, R.N.; Unocic, R.R.; Glosup, R.E.; Patterson, M.W. Fabrication and design aspects of high-temperature compact diffusion bonded heat exchangers. Nucl. Eng. Des. 2012, 249, 49–56. [Google Scholar] [CrossRef]
  14. Mortean, M.V.V.; Paiva, K.V.; Mantelli, M.B.H. Diffusion bonded cross-flow compact heat exchangers: Theoretical predictions and experiments. Int. J. Therm. Sci. 2016, 110, 285–298. [Google Scholar] [CrossRef]
  15. Hun, K.I.; Xiaoqin, Z.; Christensen, R.; Sun, X. Disign study and cost assessment of straighe, zigzag, S-shape, and OSF PCHEs for a FLiNaK-SCO2 Secondary Heat Exchanger in FHRs. Ann. Nucl. Energy 2016, 94, 129–137. [Google Scholar]
  16. Natesan, K.; Moisseytsev, A.; Majumdar, S. Preliminary issues associated with the next generation nuclear plant intermediate heat exchanger design. J. Nucl. Mater. 2009, 392, 307–315. [Google Scholar] [CrossRef]
  17. Hosseini, S.B.; Khoshkhoo, R.H.; Malabad, S.M.J. Experimental and numerical investigation on particle deposition in a compact heat exchanger. Appl. Therm. Eng. 2017, 115, 406–417. [Google Scholar] [CrossRef]
  18. Starace, G.; Fiorentino, M.; Longo, M.P.; Carluccio, E. A hybrid method for the cross flow compact heat exchangers design. Appl. Therm. Eng. 2017, 111, 1129–1142. [Google Scholar] [CrossRef]
  19. Park, M.Y.; Song, M.S.; Kim, E.S. Development of tritium permeation model for Printed Circuit Heat Exchanger. Ann. Nucl. Energy 2016, 98, 166–177. [Google Scholar] [CrossRef]
  20. Baek, S.; Kim, J.; Jeong, S.; Jung, J. Development of highly effective cryogenic printed circuit heat exchanger (PCHE) with low axial conduction. Cryogenics 2012, 52, 366–374. [Google Scholar] [CrossRef]
  21. Kim, I.H.; No, H.C. Thermal–hydraulic physical models for a printed circuit heat exchanger covering he, he–CO2 mixture, and water fluids using experimental data and cfd. Exp. Therm. Fluid Sci. 2013, 48, 213–221. [Google Scholar] [CrossRef]
  22. Lee, S.-M.; Kim, W.Y. Comparative study on performance of a zigzag printed circuit heat exchanger with various channel shapes and configurations. Int. J. Heat Mass Transf. 2013, 49, 1021–1028. [Google Scholar] [CrossRef]
  23. Lee, S.-M.; Kim, W.Y. Multi-objective optimization of arc-shaped ribs in the channels of a printed circuit heat exchanger. Int. J. Therm. Sci. 2015, 94, 1–8. [Google Scholar] [CrossRef]
  24. Mylavarapu, S.K.; Sun, X.D.; Glosup, R.E.; Christensen, R.N.; Patterson, M.W. Thermal hydraulic performance testing of printed circuit heat exchangers in a high-temperature helium test facility. Appl. Therm. Eng. 2014, 65, 605–614. [Google Scholar] [CrossRef]
  25. Ma, T.; Li, L.; Xu, X.Y.; Chen, Y.T.; Wang, Q.W. Study on local thermal-hydraulic performance and optimization of zigzag-type printed circuit heat exchanger at high temperature. Energy Convers. Manag. 2015, 104, 55–66. [Google Scholar] [CrossRef]
  26. Figley, J.; Sun, X.; Mylavarapu, S.K.; Hajek, B. Numerical study on thermal hydraulic performance of a printed circuit heat exchanger. Prog. Nucl. Energy 2013, 68, 89–96. [Google Scholar] [CrossRef]
  27. Aneesh, A.M.; Sharma, A.; Srivastava, A.; Vyas, K.N.; Chaudhuri, P. Thermal-hydraulic characteristics and performance of 3D straight channel based printed circuit heat exchanger. Appl. Therm. Eng. 2016, 98, 474–482. [Google Scholar] [CrossRef]
  28. Hamid, H.K.; Aneesh, A.M.; Autl, S.; Chaudhuri, P. Thermal hydraulic characteristics and performance of 3D wavy channel based printed circuit heat exchanger. Appl. Therm. Eng. 2015, 87, 519–528. [Google Scholar]
  29. Ngo, T.L.; Kato, Y.; Nikitin, K.; Tsuzuki, N. New printed circuit heat exchanger with S-shaped fins for hot water supplier. Exp. Therm. Fluid Sci. 2006, 30, 811–819. [Google Scholar] [CrossRef]
  30. 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]
  31. Xu, X.; Ma, T.; Li, L.; Zeng, M.; Chen, Y.; Huang, Y.; Wang, Q. Optimization of fin arrangement and channel configuration in an airfoil fin PCHE for supercritical CO2 cycle. Appl. Therm. Eng. 2014, 113, 867–875. [Google Scholar] [CrossRef]
  32. Han, C.-L.; Ren, J.-J.; Dong, W.-P.; Bi, M.-S. Numerical investigation of supercritical LNG convective heat transfer in a horizontal serpentine tube. Cryogenics 2016, 78, 1–13. [Google Scholar] [CrossRef]
  33. Bae, Y.Y. A new formulation of variable turbulent prandtl number for heat transfer to supercritical fluids. Int. J. Heat Mass Transf. 2016, 92, 792–806. [Google Scholar] [CrossRef]
  34. Jeon, S.; Baik, Y.; Byon, C.; Kim, W. Thermal performance of heterogeneous PCHE for supercritical CO2 energy cycle. Int. J. Heat Mass Transf. 2016, 102, 867–876. [Google Scholar] [CrossRef]
  35. Kruizenga, A.; Anderson, M.; Fatima, R.; Corradini, M.; Towne, A.; Ranjan, D. Heat transfer of supercritical carbon dioxide in printed circuit heat exchanger geometries. J. Therm. Sci. Eng. Appl. 2011, 3, 1–8. [Google Scholar] [CrossRef]
  36. Kim, T.H.; Kwon, J.G.; Yoon, S.H.; Park, H.S.; Kim, M.H.; Cha, J.E. Numerical analysis of air-foil shaped fin performance in printed circuit heat exchanger in a supercritical carbon dioxide power cycle. Nucl. Eng. Des. 2015, 288, 110–118. [Google Scholar] [CrossRef]
  37. Zhao, Z.; Zhang, X.; Zhao, K.; Jiang, P.; Chen, Y. Numerical investigation on heat transfer and flow characteristics of supercritical nitrogen in a straight channel of printed circuit heat exchanger. Appl. Therm. Eng. 2017, 126, 717–729. [Google Scholar] [CrossRef]
Figure 1. Thermal-physical properties of LNG at 10.5 MPa.
Figure 1. Thermal-physical properties of LNG at 10.5 MPa.
Energies 10 01828 g001
Figure 2. Error curve of linear interpolation function.
Figure 2. Error curve of linear interpolation function.
Energies 10 01828 g002
Figure 3. Schematic diagram of internal core structure of airfoil fin printed circuit heat exchanger (PCHE).
Figure 3. Schematic diagram of internal core structure of airfoil fin printed circuit heat exchanger (PCHE).
Energies 10 01828 g003
Figure 4. Schematic diagram of (a) Straight channel geometric model and (b) cross-section of straight channel.
Figure 4. Schematic diagram of (a) Straight channel geometric model and (b) cross-section of straight channel.
Energies 10 01828 g004
Figure 5. Schematic diagram of (a) the airfoil fin channel geometric model and (b) cross-section of airfoil fin channel.
Figure 5. Schematic diagram of (a) the airfoil fin channel geometric model and (b) cross-section of airfoil fin channel.
Energies 10 01828 g005
Figure 6. 3D view of computational domain for airfoil fin PCHE.
Figure 6. 3D view of computational domain for airfoil fin PCHE.
Energies 10 01828 g006
Figure 7. Schematic diagram of airfoil fin PCHE.
Figure 7. Schematic diagram of airfoil fin PCHE.
Energies 10 01828 g007
Figure 8. Dependency test of number of meshes.
Figure 8. Dependency test of number of meshes.
Energies 10 01828 g008
Figure 9. Top view of selected grid system.
Figure 9. Top view of selected grid system.
Energies 10 01828 g009
Figure 10. (a) Tested airfoil fin PCHE and (b) cold channel of airfoil fin PCHE.
Figure 10. (a) Tested airfoil fin PCHE and (b) cold channel of airfoil fin PCHE.
Energies 10 01828 g010
Figure 11. Experimental system with airfoil fin PCHE of (a) Schematic diagram of experimental set-up and (b) Photo of experimental system.
Figure 11. Experimental system with airfoil fin PCHE of (a) Schematic diagram of experimental set-up and (b) Photo of experimental system.
Energies 10 01828 g011
Figure 12. Schematic diagram for the calculation of hydraulic diameter.
Figure 12. Schematic diagram for the calculation of hydraulic diameter.
Energies 10 01828 g012
Figure 13. (a) Plots of velocity contours in airfoil fin channel and (b) straight channel when mass flux is 325 kg/m2·s.
Figure 13. (a) Plots of velocity contours in airfoil fin channel and (b) straight channel when mass flux is 325 kg/m2·s.
Energies 10 01828 g013
Figure 14. (a) Heat transfer coefficient h and (b) Nusselt number as a function of mass flux.
Figure 14. (a) Heat transfer coefficient h and (b) Nusselt number as a function of mass flux.
Energies 10 01828 g014
Figure 15. (a) Euler number and (b) Fanning friction factor as a function of mass flux.
Figure 15. (a) Euler number and (b) Fanning friction factor as a function of mass flux.
Energies 10 01828 g015
Figure 16. Nu/Eu as a function of mass flux.
Figure 16. Nu/Eu as a function of mass flux.
Energies 10 01828 g016
Figure 17. Velocity contour in PCHE with airfoil fins arranged at Lv = 1.67 mm.
Figure 17. Velocity contour in PCHE with airfoil fins arranged at Lv = 1.67 mm.
Energies 10 01828 g017
Figure 18. Effect of Lv on (a) Nusselt number, (b) pressure drop, (c) Euler number and (d) Nu/Eu.
Figure 18. Effect of Lv on (a) Nusselt number, (b) pressure drop, (c) Euler number and (d) Nu/Eu.
Energies 10 01828 g018
Figure 19. Velocity contour in PCHE with airfoil fins arranged at Ls = 4 mm.
Figure 19. Velocity contour in PCHE with airfoil fins arranged at Ls = 4 mm.
Energies 10 01828 g019
Figure 20. Effect of Ls on (a) heat transfer coefficient h and (b) Nusselt number.
Figure 20. Effect of Ls on (a) heat transfer coefficient h and (b) Nusselt number.
Energies 10 01828 g020
Figure 21. Effect of Ls on (a) Euler number and (b) Nu/Eu.
Figure 21. Effect of Ls on (a) Euler number and (b) Nu/Eu.
Energies 10 01828 g021
Table 1. Property correlations of LNG at 10.5 Mpa.
Table 1. Property correlations of LNG at 10.5 Mpa.
(1) 121–227 K
ρ = 8256.5933 − 285.3922 T + 4.3559 T2 − 0.035267 T3 + 1.5894 × 10−4 T4 − 3.7784 × 10−7 T5 + 3.6917 × 10−1 T6
Cp = −1.0199 × 106 + 3.8183 × 104 T − 589.1061 T2 + 4.8079 T3 − 0.02189 T4 + 5.28734 × 10−5 T5 − 5.23296 × 10−8 T6
λ = 1.6663 − 0.051726 T + 7.9379 × 10−4 T2 − 6.6085 × 10−6 T3 + 3.0653 × 10−8 T4 − 7.4993 × 10−11 T5 + 7.5617 × 10−14 T6
μ = 4.0977 × 10−3 − 1.2117 × 10−4 T + 1.5988 × 10−6 T2 − 1.1623 × 10−8 T3 + 4.8332 × 10−11 T4 − 1.0811 × 10−13 T5 + 1.01103 × 10−16 T6
(2) 227–315 K
ρ = −3.89476 + 7.7791 × 103 T − 63.6984 T2 + 0.273703 T3 − 6.4991 × 10−4 T4 + 8.06233 × 10−7 T5 − 4.06261 × 10−10 T6
Cp = −4.48732 × 107 + 9.7441 × 105 T − 8.78917 × 103 T2 + 42.17166 T3 − 0.011355 T4 + 1.62727 × 10−4 T5 − 9.6971 × 10−8 T6
λ = −99.92683 + 2.24495 T − 0.0208241 T2 + 1.02288 × 10−4 T3 − 2.8093 × 10−7 T4 + 4.09353 × 10− 10 T5 − 2.47389 × 10−13 T6
μ = 1.58763 × 10−2 + 3.9445 × 10−4 T − 3.96736 × 10−6 T2 + 2.08564 × 10−8 T3 − 6.07355 × 10−11 T4 + 9.3195 × 10−14 T5 − 5.8999 × 10−17 T6
(3) 315–385 K
ρ = 9403.1676 − 140.8996 T + 0.90485 T2 − 3.14226 × 10−3 T3 + 6.19397 × 10−6 T4 − 6.55485 × 10−9 T5 + 2.905 × 10−12 T6
Cp = −4.28439 × 105 − 6.56712 × 103 T + 42.6554 T2 − 0.14901 T3 + 2.94672 × 10−4 T4 − 3.12279 × 10−7 T5 + 1.383995 × 10−10 T6
λ = 2.2909 − 0.03353 T + 2.35313 × 10−4 T2 − 8.3695 × 10−7 T3 + 1.68679 × 10−9 T4 − 1.8217 × 10−12 T5 + 8.22556 × 10−16 T6
μ = 4.92979 × 10−3 − 8.37086 × 10−5 T + 5.94466 × 10−7 T2 − 2.25336 × 10−9 T3 + 4.8073 × 10−12 T4 − 5.47099 × 10−15 T5 + 2.5942 × 10−18 T6
Table 2. Comparisons of simulation and experiment results.
Table 2. Comparisons of simulation and experiment results.
Pressure (MPa)Experiment Results of ∆P/L (Pa/m)Simulation Results of ∆P/L (Pa/m)Error (%)Experiment Results of Tout (K)Simulation Results of Tout (K)Error (%)
5.537,090.6117235,226.125.03%279.45278.130.472%
633,988.773732,023.75.78%281.15281.960.288%
6.525,650.6762427,635.157.74%282.55283.671.12%
724,175.7466926,985.3711.62%284.35285.960.669%
7.523,834.3093624,865.354.33%285.65288.871.127%

Share and Cite

MDPI and ACS Style

Zhao, Z.; Zhao, K.; Jia, D.; Jiang, P.; Shen, R. Numerical Investigation on the Flow and Heat Transfer Characteristics of Supercritical Liquefied Natural Gas in an Airfoil Fin Printed Circuit Heat Exchanger. Energies 2017, 10, 1828. https://doi.org/10.3390/en10111828

AMA Style

Zhao Z, Zhao K, Jia D, Jiang P, Shen R. Numerical Investigation on the Flow and Heat Transfer Characteristics of Supercritical Liquefied Natural Gas in an Airfoil Fin Printed Circuit Heat Exchanger. Energies. 2017; 10(11):1828. https://doi.org/10.3390/en10111828

Chicago/Turabian Style

Zhao, Zhongchao, Kai Zhao, Dandan Jia, Pengpeng Jiang, and Rendong Shen. 2017. "Numerical Investigation on the Flow and Heat Transfer Characteristics of Supercritical Liquefied Natural Gas in an Airfoil Fin Printed Circuit Heat Exchanger" Energies 10, no. 11: 1828. https://doi.org/10.3390/en10111828

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