abstract - bura.brunel.ac.uk€¦  · web viewa clear increase in the earth’s average...

44
PERFORMANCE EVALUATION OF A MULTI-PASS AIR-TO-WATER THERMOSYPHON- BASED HEAT EXCHANGER H. Mroue 1,a , J.B. Ramos 2 , L.C. Wrobel 3,a , H. Jouhara *,1,a 1 Institute of Energy Futures, RCUK Centre for Sustainable Energy Use in Food Chains (CSEF), 2 Faculty of Computing, Engineering and Science, University of South Wales, Pontypridd CF37 1DL, United Kingdom 3 Institute of Materials and Manufacturing a College of Engineering, Design and Physical Sciences, Brunel University London, Uxbridge UB8 3PH, UK * Corresponding author: Tel. +44 1895 267805; Email: [email protected] Keywords: heat pipe, thermosyphon, heat exchanger, CFD, effectiveness Abstract The project reported in this paper used CFD as a tool to investigate the effect of multi-pass on the shell side heat transfer of a heat exchanger system. The heat exchanger system is equipped with six vertical thermosyphons transferring thermal energy from a heat source (air) to a heat sink (water). The CFD model has been experimentally validated. The two-phase change processes inside the thermosyphons were not modelled during the simulation. Instead, the thermosyphons were treated as solid rods with a constant thermal conductivity, which was calculated theoretically by applying the thermal resistance analogy with the aid of convection, boiling and condensation correlations found in the literature. The heat source consists of multiple air passes on the evaporator section of the thermosyphons and two water passes on the condenser section. Three different arrangements on the evaporator section were investigated with one, two or three shell passes and the thermal performance compared for the three configurations. The investigation was performed at various inlet conditions: a range of air inlet temperatures (100, 150, 200 and 250°C) and mass flow rates (0.05, 0.08, 0.11 and 0.14 kg/s). The water inlet conditions were kept constant (a temperature of 15°C and a mass flow rate of 0.08 kg/s). The overall rate of heat transfer was obtained by both CFD and a theoretical model, and the results lay within 15% of the experimental data. The numerical 1

Upload: ngonhu

Post on 09-Sep-2018

212 views

Category:

Documents


0 download

TRANSCRIPT

PERFORMANCE EVALUATION OF A MULTI-PASS AIR-TO-WATER THERMOSYPHON-BASED HEAT EXCHANGER

H. Mroue1,a, J.B. Ramos2, L.C. Wrobel3,a, H. Jouhara*,1,a

1Institute of Energy Futures, RCUK Centre for Sustainable Energy Use in Food Chains (CSEF),

2Faculty of Computing, Engineering and Science, University of South Wales, Pontypridd CF37 1DL, United Kingdom3Institute of Materials and Manufacturing

aCollege of Engineering, Design and Physical Sciences, Brunel University London, Uxbridge UB8 3PH, UK*Corresponding author: Tel. +44 1895 267805; Email: [email protected]

Keywords: heat pipe, thermosyphon, heat exchanger, CFD, effectiveness

Abstract

The project reported in this paper used CFD as a tool to investigate the effect of multi-pass on the shell side heat transfer of a heat exchanger system. The heat exchanger system is equipped with six vertical thermosyphons transferring thermal energy from a heat source (air) to a heat sink (water). The CFD model has been experimentally validated. The two-phase change processes inside the thermosyphons were not modelled during the simulation. Instead, the thermosyphons were treated as solid rods with a constant thermal conductivity, which was calculated theoretically by applying the thermal resistance analogy with the aid of convection, boiling and condensation correlations found in the literature. The heat source consists of multiple air passes on the evaporator section of the thermosyphons and two water passes on the condenser section. Three different arrangements on the evaporator section were investigated with one, two or three shell passes and the thermal performance compared for the three configurations. The investigation was performed at various inlet conditions: a range of air inlet temperatures (100, 150, 200 and 250°C) and mass flow rates (0.05, 0.08, 0.11 and 0.14 kg/s). The water inlet conditions were kept constant (a temperature of 15°C and a mass flow rate of 0.08 kg/s). The overall rate of heat transfer was obtained by both CFD and a theoretical model, and the results lay within 15% of the experimental data. The numerical predictions demonstrated that the k−εRealizable turbulence model is a reliable tool for predicting heat transfer and fluid flow in such heat exchangers.

1. Introduction

It is generally accepted that throughout history, especially after the industrial revolution, humans have had a negative impact on the environment. A clear increase in the earth’s average temperature has been observed for some years and the world’s leading climate scientists believe that this rise in temperature is directly related to mankind’s activities [1], such as the burning of fossil fuels, deforestation and livestock farming. Those activities produce gases which act in a similar way to the glass in a greenhouse, permitting short-wavelength solar radiation to be incident on the earth’s surface but absorbing long-wavelength infra-red radiation from the earth, thereby increasing global temperatures. Gases produced by human activities include, in particular, carbon dioxide, methane, nitrous oxide and fluorinated gases. Carbon dioxide is the major greenhouse gas (GHG), which contributes 64% of man-made global warming and its concentration in the atmosphere is currently 40% higher than when industrialisation began [2]. Other GHGs contribute less to global warming: 17% for methane and 6% for nitrous oxide, although they are more efficient at absorbing infra-red radiation than carbon dioxide.

1

Average global temperatures have been observed to have risen by about 0.85°C in the past 150 years (50% of the rise in the past 20 years) and they are still subject to further increases. If they exceed 2°C, there is a risk of dangerous changes in human and natural systems [2] and actions have been taken by the United Nations Framework Convention on Climate Change (UNFCCC). The objective was to level off GHG emissions in this decade and to reduce them by 50% compared to 1990 levels by 2050. In order to achieve a reduction in GHGs, the European Union (EU) has imposed some policies including the increased use of renewable energy sources and a continual improvement in the energy efficiency of a wide range of systems [3]. EU leaders have also set several targets to be met by 2020. These targets were proposed in 2007 and agreed in 2009, including a 20% reduction in GHG emissions, 20% of energy to come from renewables and a 20% improvement in energy efficiency.

Companies involved in building services and process industries have therefore been forced to design more sustainable and energy efficient systems [4–6] to meet the EU targets. Exhaust gases generated from such activities release both GHGs and waste heat to the atmosphere and they are a significant contributor to global warming. Waste heat could be recovered and/or recycled through heat exchanger (HX) systems to be reused within the industrial processes, which would save energy and decrease the power consumption coming from fossil fuels, hence reducing GHG emissions.

In this paper, a heat exchanger system designed to recover waste heat using heat pipe (HP) technology has been investigated. The shell side heat transfer will be investigated theoretically, experimentally and numerically for multiple shell side fluid passages.

2. Heat pipes

To tackle the energy problem caused by the rapid decrease of oil and fossil fuel suppliers as mentioned by Olabi [7,8], more research and developments are required in the energy sector. One such recent development involves the use of heat pipe based heat exchanger systems in waste heat recovery. Heat pipes are devices that facilitate the transfer of thermal energy between two media (heat source and heat sink) within the heat exchanger. They are also described as superconductors because of the high heat transfer capability over large distances using a small temperature difference [9]. A heat pipe is a device that contains a small amount of liquid known as the working fluid, in an evacuated sealed tube. The common cross-sectional area of the tube is either cylindrical or rectangular. However, the shapes of the cross-section and design are not limited to these two options and can be altered, depending on the application [10].

A heat pipe may be divided in three different areas: evaporator, adiabatic section and condenser. The working fluid is located in the evaporator section where heat is applied on the external wall of the pipe. Heat is then absorbed by the working fluid through the evaporator wall, essentially in the form of latent heat. When the temperature of the liquid pool exceeds the saturation temperature, nucleation sites and small bubbles start to form on the inner wall of the evaporator. The absorption of more heat causes the bubbles to grow in size and the vapour generated travels towards the condenser under the effect of the vapour pressure difference between the evaporator and condenser section. Before reaching the condenser, the vapour travels through the adiabatic section where heat is conserved. Upon reaching the condenser, the vapour contacts the condenser inner wall which is at a temperature below saturation. The vapour will then condense back into liquid, forming a thin layer of liquid film, flowing back along the walls to the evaporator and hence completing the cycle. The conventional heat pipe returns the liquid condensate by capillary forces induced by the presence of a wick structure; however, a thermosyphon (TS), also known as wickless heat pipe (Figure 1 Thermosyphon schematic.), relies upon gravity to return the liquid condensate back to the evaporator section.

2

The heat exchanger in this paper is equipped with six thermosyphons. The major limitation in the operation of a thermosyphon compared to a conventional heat pipe is the orientation of the device, where the condenser must be located above the evaporator for the device to operate. Many factors can affect the thermal performance of a thermosyphon including filling ratio, working fluid, inclination angle, aspect ratio, working temperature and the length of each section. Negishi and Sawada [11] experimentally investigated the thermal performance of a thermosyphon at different inclination angles and filling ratios. The investigation was conducted with two different working fluids, water and ethanol. For water, it was found that the optimum heat transfer rate was achieved for an inclination between 20 and 40° from the horizontal and a filling ratio (FR) of between 25 and 60%, whereas for ethanol an angle greater than 5° was needed and a filling ratio of between 40% and 75%. An extra 5 to 10% of the working fluid is recommended in the thermosyphon to avoid the break-down of the liquid film during operation [12]. Noie et al. [13] conducted an experimental investigation of a copper thermosyphon with distilled water as the working fluid. A combination of inclinations between 5 and 90° were tested with filling ratios of 15, 22 and 30%. The thermosyphon was seen to perform best for angles between 15 and 60°. The “Geyser” boiling phenomenon was observed for filling ratios above 30% and this should be avoided as it could damage the top end of the condenser from continuous strikes by liquid slugs. Jouhara et al. [14] conducted an experimental investigation of a wickless heat pipe made of copper with two working fluids: water and ethanol-water azeotrope. The condenser was 12° inclined from the evaporator section. The thermosyphon was able to operate at evaporator inclination angles ranging from 0 to 90°, including the horizontal position.

Heat pipes and thermosyphons have become popular because they can be integrated into heat exchanger systems. Such systems have been applied in various industries, including power generation, electronic cooling systems, waste heat recovery, air conditioning systems and food industries [10,13]. The use of heat pipes improves the thermal performance of systems by transporting a large amount of heat through a small cross-sectional area. Jouhara and Merchant [15] carried out an experimental investigation of an air-to-air thermosyphon based heat exchanger (TSHX) used in building services. The cross-flow heat exchanger was equipped with nine thermosyphons in an in-line arrangement. The thermosyphons were filled with water as the working fluid and had fins installed on both the evaporator and condenser sections to enhance the convective heat transfer. The heat load and inclination angle were varied. The thermal resistance and the overall effectiveness were found to vary with the air mass flow rate. A prediction of the overall effectiveness was achieved through the variation of the working temperature of the pipes. A theoretical model of the overall effectiveness was also built using mathematical correlations. Danielewicz et al. [5] conducted an experimental and analytical investigation of an air-to-air TSHX where a total of 100 thermosyphons were installed in 10 rows. The thermosyphons were 2.2 m in length, made of carbon steel and methanol was used as the working fluid. Aluminium fins were also installed along the length. An effectiveness NTU (number of transfer units) method was developed in [5] and experimentally validated to predict variables such as the overall heat transfer coefficient, pressure drop and heat transfer duty. Good agreement was observed between the experiment and theoretical predictions. The prediction model developed can be used for future designs of typical cross-flow air-to-air TSHX systems.

Noie- Baghban et al. [16] conducted an experimental and theoretical investigation of an air-to-air cross flow HPHX that has applications in operating theatres in hospitals. From the experimental results obtained, Noie-Baghban et al. [16] noted that the following actions would improve the efficiency of an HPHX system:

- Installing fins around the heat pipes- Increasing the number of rows of heat pipes- Minimizing heat loss- Applying an appropriate method of charging the heat pipes to prevent accumulation of non-

condensable gases

3

- Perfect sealing of the heat pipes to avoid a loss of vacuum

Due to the complex fluid mechanics and heat transfer mechanisms present in heat exchanger systems, CFD has become an alternative for researchers studying the flow and heat transfer behaviour for multi-pass heat exchangers [17,18] and inside heat pipes [19–21]. In addition, flow patterns and temperature fields can be obtained which experimental testing cannot achieve [22]. Ramos et al. [17] conducted both CFD and an experimental investigation on a similar TSHX system where there was only one air pass on the evaporator side. The thermal network resistance was also applied alongside the effectiveness method to model the thermal performance of the system. In this paper, the effect of changing the number of fluid passes on the evaporator section of a TSHX system and the shell side heat transfer will be investigated theoretically, numerically and experimentally. Such investigations are not well covered in the literature.

3. Experimental apparatus

The heat exchanger under investigation consisted of six thermosyphons installed vertically and arranged in two rows, each with three thermosyphons. The thermosyphon bundle was in a staggered arrangement on the evaporator (air) side and in an in-line arrangement on the condenser (water) side. The TSHX unit was manufactured by Econotherm Ltd. (Bridgend, Wales), a company which specialises in manufacturing waste heat recovery systems integrated with heat pipe technology. The unit under study was a small scale version of a real shell-and-tube heat exchanger with a Disk-and-Doughnut baffle arrangement currently on the market. The thermosyphons were made of carbon steel and the surfaces were chemically treated from the inside in order to prevent corrosion or any type of chemical interaction with the working fluid (water) inside the pipes, where continuous boiling and condensation processes occur during operation. During the charging process, the inner pipe wall surface was superheated by steam to achieve oxidation of the carbon steel [23]. Such processes lead to a successful operation and prolonged life of the test unit.

Figure 2 shows the unit under investigation. The TSHX was 2 m in length and each air pass covered 600 mm of the total length of the thermosyphons. A maximum of three air passes could be installed on the shell-side of the heat exchanger, giving a total length of 1.8 m for the evaporator and 200 mm for the condenser section. The thermosyphons had the same length as the TSHX and had an outside diameter of 28 mm and a wall thickness of 2.5 mm. The same amount of working fluid (water) was injected into each thermosyphon during the charging process, filling each thermosyphon to a height of 600 mm. Three different configurations of the test unit were investigated as follows:

Case 1, one air pass Case 2, two air passes Case 3, three air passes

The thermosyphon condenser sections were installed in a U-shaped flow passage (Figure 3).

Figure 4 shows the three different configurations of the test unit where the air flow in each case is indicated by a black arrow. The TSHX under investigation was well insulated to minimise heat loss in order to achieve, to a reasonable extent, the boundary conditions assumed during the numerical modelling, where external walls were taken to be 100% adiabatic.

4. Operational procedure

The TSHX in this paper operated between two different fluid circuits, namely a hot air circuit and a cold water circuit. In Figure 5, the hot side is represented by solid lines while the cold side in shown by dashed lines. The air flow was generated by a variable frequency fan and four different flow rates were considered:

4

0.05, 0.08, 0.11 and 0.14 kg/s. The fan blew the air into the heater, where its temperature was controlled by a thermocouple feedback loop. The temperature of the hot air entering the HX was varied between 100 and 250°C in steps of 50°C. The air was then forced to enter the shell-side first air pass, flowing across the staggered tube bundle where heat was absorbed by the thermosyphons. Passes were connected by elbows, directing the air flow externally in U-shaped passages, with the hot air crossing the tube bundle once in each pass. Finally, the hot air exited the HX and returned to the fan in a closed loop circuit.

Water was the fluid medium on the condenser side used to recover the heat absorbed by the evaporator section. Water was initially located in a tank at an ambient temperature of approximately 15°C and then pumped through a flow meter before flowing at a rate of 0.08 kg/s around the thermosyphons, one by one in an in-line configuration (Figure 3 and Figure 5).

A total of 22 K-type thermocouples were installed at various locations across the test unit, at the inlet and outlet of each air pass and at the inlet and outlet of the cooling water (Figure 6). Thermocouples were also brazed on the outer surface of the thermosyphons in each air pass.

5. Theoretical Modelling

The thermal resistance network approach was applied in order to evaluate the performance of the whole TSHX system, as it has previously given satisfactory results as described by many researchers [17,18,24]. This method divides the TSHX system into different components where the various heat transfer modes can be represented, namely conduction, convection, condensation and boiling heat transfer.

Convective heat transfer coefficient (Air-Wall, Water-Wall)

The thermal resistance of the entire TSHX system is outlined in Figure 7 and Figure 8 for the three configurations. Starting from the hot air entering the heat exchanger, heat is transferred by convection to the pipe’s outer surface in the first air pass. The same happens with the second and third air passes. The resistance to convective heat transfer between the shell side fluid (air or water) and the pipe outer surface is listed in Table 1: equations (11) and (16). The convective contact area Ae at the evaporator is the same for each air pass due to the same height of the section (0.6m). However, the area at the condenser ( Ac) is smaller due to the shorter height (0.2 m).

Conduction through the thermosyphon wall

The heat is transferred radially across the thermosyphon wall by conduction where the resistance is calculated using equation (12). Axial conduction along the length of the thermosyphon was not taken into account due to its negligible value.

Pool boiling heat transfer

5

In case 1, where one air pass is applied across the six thermosyphons, a filling ratio of 100% is considered as the evaporator is completely full with water (Figure 8). Therefore, the only mode of heat transfer that exists inside the evaporator is nucleate pool boiling where the correlation of Rohsenow can be applied and shown by equation (13).

Nucleate film boiling

In cases 2 and 3 (Figure 7 and Figure 8), the evaporator length was increased by 0.6 m in each case, keeping the same amount of working fluid in each thermosyphon. The filling ratios in cases 2 and 3 are therefore 1/2 and 1/3 , respectively. An extra mode of convective heat transfer is considered within the thermosyphon in the second and the third air passes. This mode of heat transfer takes place through the liquid condensate film returning to the evaporator pool and is referred to as nucleate film boiling [18], equation (14).

Film condensation

The vapour inside the thermosyphon travels vertically towards the condenser section where, upon coming into contact with the inner pipe cold surface, it condenses back to liquid, releasing its latent heat, which will be absorbed by water surrounding the six thermosyphons. This mode of heat transfer is essential for a thermosyphon to complete its thermodynamic cycle. The heat transfer coefficient is calculated through an expression derived by Nusselt [17,18], equation (15).

Thermal resistance and conductivity k

The thermal resistance in equation (1) is defined as the ratio of the temperature difference across a component to the corresponding rate of heat transfer [25]. The application of this approach would identify which component within the HX system restricts the majority of the heat transfer and the design could be revised.

Ri=∆ T i

Q i(1)

In the first case and referring to Figure 8, the internal resistance of one thermosyphon is calculated by the following equation:

RTS=( Rk ,Wall+Rh ,NPB )pass1+( Rh , NFC+R k ,Wall )Condenser (2)

For the second and third cases, the internal resistance of one thermosyphon is presented in equations (3) and (4), respectively:

RTS=(( 1Rk , Wall+Rh , NPB )pass1

+( 1Rk , Wall+Rh , NFB )pass2)

−1

+( Rh , NFC+Rk ,Wall )Condenser (3)

6

RTS=(( 1Rk , Wall+Rh , NPB )pass1

+( 1Rk , Wall+Rh , NFB )pass2

+( 1Rk , Wall+Rh , NFB )pass3)

−1

+( Rh , NFC+Rk , Wall)Condenser(4)

Assuming all the six thermosyphons have equal internal resistances, the equation can be simplified as follows:

1R6 TPs

=∑i=1

6 1R i

(5)

Adding more thermosyphons to the heat exchanger system will add extra resistances in a parallel configuration as can be seen from equation (5). Therefore, the internal and total resistance of the whole system will be reduced and more heat will be recovered,

R6TS=RTS

6(6)

The total thermal resistance RTotal of the system is needed to calculate the overall heat transfer by adding the external convective heat transfer, giving a total thermal resistance for case 1 as follows:

RTotal=R6TS+Rh , Air ,Wall+Rh ,Water ,Wall (7)

For the second the third cases, the resistance of the convective heat transfer in the second and third pass from the air to the pipe outer wall will also be added.

The internal thermal resistance of a single thermosyphon (RTS) is an essential parameter in this paper for the purpose of calculating the effective thermal conductivity of a thermosyphon. Later in the paper, using ANSYS Fluent, the calculated thermal conductivity will be used as a boundary condition where it replaces the phase change heat transfer mechanisms inside the thermosyphon by pure conduction heat transfer,

RTS=L

k ATS∴k= L

RTS ATS(8)

Equation (8) describes the parameters that relate the thermal conductivity k to the thermal resistance RTS, where L is the length of the thermosyphon and ATS its cross-sectional area.

6. Numerical Modelling (ANSYS Fluent)

The ANSYS Fluent CFD modelling package was used to model the TSHX as a full three-dimensional system. As mentioned above, the thermosyphons were represented by six solid rods with constant conductivity, k , calculated theoretically using correlations found in the literature. The numerical tool presented in this paper is used to characterise and assess the thermal performance of a real TSHX which can

7

be used to design and optimise future heat exchangers. The TSHX was sketched in ANSYS workbench 16.2 for the three different configurations (cases 1, 2 and 3).

Grid analysis

A grid analysis study was initially conducted in order to choose an optimised mesh to run the three different cases at various air inlet conditions. The analysis compared the results with experimental data to ensure an adequate number of elements while minimising computational cost and time. Simulations were run at different test conditions and mesh densities. Three different mesh densities were chosen: coarse, medium and fine. The comparison between experimental and numerical results mainly focused on the evaporator and condenser outlet temperatures, i.e. air and water outlet temperatures. For comparison, one mesh density was chosen and applied to the three test cases and this was based on a grid analysis for the most complicated case, i.e. case 3.

As mentioned above, mesh refinement was conducted on the most complicated case (case 3) by changing the relevance value in the global mesh sizing control. Relevance is a useful parameter to achieve an automatic global refinement or coarsening of the mesh. The default value of the relevance is 0 and the value can be changed between -100 to 100, where the values of -100 and 100 correspond respectively to a very coarse mesh and a very fine mesh. Three different mesh densities were chosen based on increasing the relevance value from 0 (coarse) to 100 (fine) where a relevance of 50 was considered as a medium size mesh. Table 2 shows the different parameters which were taken into account when analysing the mesh size and quality. It can be clearly noted that an increase of the relevance value increases the number of elements and improves the quality of the mesh by reducing the skewness of the cells which can cause convergence issues during the simulation. The mesh generated was dominated by tetrahedral shaped elements located everywhere in the control volume except in the elbows and solid rods, which were dominated by hexahedral elements (Figure 9 and Figure 10).

In order to simulate the convective heat transfer between the fluids (air and water) and the solid rods accurately, an inflation layer consisting of three layers was placed around the six solid rods (Figure 10). The presence of an inflation layer would increase the skewness of the mesh as it may cause a stair-step in the mesh. However, it leads to more accurate results as the critical regions in the domain are the solid-fluid interfaces, therefore a thin layer is needed to develop a smooth thermal boundary layer.

Figure 11 to Figure 14 show the percentage error between CFD and experimental results for case 3 at two different air inlet conditions (150 and 250°C). The figures clearly show that the medium mesh and the fine mesh produce almost identical results, which demonstrates that increasing the number of elements beyond the medium mesh will just increase computational time and cost with no significant improvement in the accuracy of the results. Hence, the fine mesh was excluded from the analysis. The figures also illustrate that the coarse mesh has performed better, in terms of producing results closer to the experimental results, than the medium mesh. The coarse mesh, however, caused issues regarding convergence as the residuals showed a significant fluctuation for all the transport equations including continuity, energy and momentum. Based on the previous analysis, the medium mesh was chosen for all three cases for the rest of the simulations. Another characteristic can be deduced from the above results related to the signs of the errors. The positive values for air outlet temperatures indicate that the CFD over-predicts the air outlet temperatures. The heat

8

exchanger wall was set to be 100% adiabatic during the simulations, which neglects any heat loss across the boundary, resulting in slightly elevated air outlet temperatures.

Fluent boundary conditions (BC)

The model within the heat exchanger can be described as a low speed single phase flow with a very small density change, therefore a pressure-based solver was chosen. This produces better results and is more efficient in steady-state simulations.

shows the different boundary types and test conditions that were chosen for the simulation. The mass flow rate was defined at both evaporator and condenser inlets, while the pressure was defined at both outlets.

On the evaporator (air) side, four different mass flow rates were considered: 0.05, 0.08, 0.11 and 0.14 kg/s. For each mass flow rate, four different inlet temperatures were examined: 100, 150, 200 and 250°C, giving a total of 16 different combinations of mass flow rate and inlet temperature. On the condenser (water) side, only one mass flow rate and one inlet temperature were chosen, 0.08 kg/s and 15°C respectively. In Fluent, the heat exchanger walls were set as adiabatic (Q̇=0). In the case of external flows, it is recommended that turbulence intensity and viscosity ratio should be specified at the inlet [26]. No information was available for the inlet turbulence intensity of either air or water flows, and therefore the default value of 5% was chosen for both inlets.

The Realizable k−ε turbulence model was adopted in the simulations as it is suitable for standard cases where a model involves moderate swirls, vortices and flow separation without providing an accurate resolution of the boundary layer separation [27]. The residuals of all the transport equations were set to 1e-06 to give enough time for the simulation to reach a steady-state condition. In the solution methods, the SIMPLE algorithm scheme was chosen with satisfactory results.

7. Results and discussion

Comparison of results (CFD, theoretical and experimental)

Figures 15 to 20 compare experimental results for the three cases at two different evaporator inlet temperatures, 150°C and 250°C. The figures show the rate of heat transfer for each pass where their sum represents the overall heat transfer. An increase in the air inlet temperature resulted in increases in the temperature differences across the evaporator and the condenser, and therefore an increase in the overall rate of heat transfer. The heat transfer also increases as the air mass flow rate is increased. For a constant air inlet temperature, the heat duty across the first pass (blue area) is approximately the same for the three cases where only pool boiling is taking place, and similarly when comparing the heat transfer across the second pass (red area) for the same evaporator inlet temperature. However, the heat transfer mode occurring in the second pass is nucleate film boiling.

According to the theoretical calculations shown in Figure 15 to Figure 20, pool boiling represented 100% of the evaporator heat input for case 1, 64% for case 2 and 53% for case 3 at a given inlet temperature (150 and 250°C). The same trend was also observed for the remaining air inlet temperatures (100 and 200°C). For a given air inlet temperature the addition of air passes led to more conductive paths in parallel, hence a lower overall resistance and higher overall heat transfer.

9

Figure 21 to Figure 23 compare the theoretical results, CFD and experimental data at different air inlet temperatures and mass flow rates where the lowest and highest heat transfer rates occur at the following combinations of air inlet conditions: (0.05kg/s, 100°C) and (0.14kg/s, 250°C), respectively. Not surprisingly, the highest rate of heat transfer occurred with the highest values of air inlet parameters for case 3, while the lowest rate of heat transfer occurred at the lowest values of these parameters for case 1.

As stated above, a theoretical model of all cases was developed using existing correlations available in the literature and these have been previously used by many researchers, including Ramos et al. [17] and Mroue et al. [18], and shown to be reasonably accurate. Correlations from Zhukauskas, Rohsenow and Nusselt [17,18] were used to model the convection, boiling and condensation heat transfer mechanisms inside and outside the thermosyphons. The overall heat transfer rate was then calculated and compared with experimental and numerical (CFD) results. It can be clearly seen that the majority of the CFD and theoretical results lie within a 15% envelope of the experimental heat transfer.

Looking at Figure 21 to Figure 23, the CFD and theoretical heat transfer rates closely follow the experimental line (black) at low values of air inlet parameters. The percentage error increases at higher values. This can be due mainly to the lack of information on turbulence, which is needed in Fluent at air and water flow inlets for accurate simulation. As mentioned before, the turbulence intensity of both air and water were kept at the 5% default value. However, increasing the mass flow rate or the inlet air temperature will lead to higher turbulence and, therefore, higher convective heat transfer which was not taken into account in the simulation.

Figure 24 and Figure 25 show the percentage error between experimental and CFD results in the outlet temperature on the evaporator and condenser sections at different inlet conditions. For the three cases, the figures have shown that the majority of the results lie within a 15% envelope as mentioned before.

Another comparison has been made between CFD and experimental results on the outlet temperature of both evaporator and condenser. The comparison is illustrated from Figure 26 to Figure 31. The line y=x represents a perfect correlation between both CFD and experimental results. The CFD results show a slight overestimation of the outlet temperature of the evaporator, but this was expected as the CFD imposes a 100% adiabatic wall whereas, in reality, complete insulation is hardly achieved.

CFD contours

Contours of the temperature distribution in the evaporator section are displayed in Figure 32. All cases were run at an air inlet temperature of 150°C and a mass flow rate of 0.14 kg/s. It can be clearly seen how the air temperature gradually decreases as the air flows towards the evaporator outlet, with the drop in temperature and hence the rate of heat transfer increasing as more passes are added. It can also be seen that the air

10

temperature distribution within connecting ducts is essentially uniform due to the imposed adiabatic wall boundary condition.

Figure 33 shows the magnitude of the velocity vectors in the evaporator section for the three cases at 150°C air inlet temperature and 0.14 kg/s mass flow rate. The behaviour of the velocity profile in the three cases is similar. The blue colour refers to a zero mean velocity where the air flow is trapped in a recirculation zone outside the flow main stream, mainly at top and bottom sections of each pass. The red colour corresponds to the highest velocity the air flow has reached; it can be seen that this occurs only in specific sections of the HX, more specifically at the bends of the elbows where the air flow gains more speed due to the change in direction. This plot is helpful in identifying the wake regions within the heat exchanger, which can help in the optimisation of the HX design.

A section of the condenser is illustrated in Figure 34 where the temperature variation is shown across the heat pipes and water. The cross-section has been taken in the middle of the condenser, therefore at a height of 1.9 m from the bottom of the evaporator. The simulation was run at a water inlet temperature of 15°C, the cross-section was higher than the water inlet, which explains the minimum value of 17°C in the scale. The figure clearly shows how the water absorbs heat from the pipes as it flows around them in a U-shaped duct as the temperature of the pipes gradually increases. The pipe at the water inlet has the lowest temperature, indicating the highest convective heat transfer since the temperature difference between the pipe and water is the highest here. The water reaches the pipe at the outlet with the smallest temperature difference, therefore the convective heat transfer is least here. Comparing the pipe temperatures for the three cases, it is obvious that the three passes configuration has the highest temperature difference for water between inlet and outlet, and hence the highest heat transfer rate.

8. Effectiveness

A numerical effectiveness-NTU (ε−NTU ¿model was developed on the whole heat exchanger in order to analyse the effect of different inlet conditions and air passes on the rate of heat transfer. Jouhara and Merchant [15] and Ramos et al. [17] developed a similar model where the heat exchanger was treated as two separate heat exchangers consisting of the evaporator and condenser coupled by the thermosyphons as a mode of heat transfer.

Figure 35 shows the variation of the overall effectiveness for the different inlet conditions for Case 3. A consistent downward trend can be observed for the different inlet conditions, this trend being in good agreement with Jouhara and Merchant [15] and Ramos et al. [17] where a higher effectiveness is achieved by increasing the evaporator inlet temperature while reducing its mass flow rate. Figure 36 and Figure 37 show the change of the effectiveness with the number of transfer units (NTU) at different inlet conditions for the three test cases, for air inlet temperatures of 100°C and 250°C, respectively. Each case is plotted at four different air inlet flow rates (0.05, 0.08, 0.11 and 0.14 kg/s). It can be clearly seen that more air passes installed at the evaporator result in a higher effectiveness of the heat exchanger, although the impact on pressure drop should always be taken into account. Both figures show a linear increase in effectiveness with NTU, which agrees with the effectiveness-NTU graphs plotted by Incropera and DeWitt [28] over these ranges of effectiveness and NTU.

11

9. Error Analysis

The thermocouples were probably the main source of error in the experiment conducted. In this section, an uncertainty study will be carried out to investigate the error propagation on the heat recovered by the condenser at the two extreme air inlet temperatures (100°C and 250°C) for the three configurations. The uncertainty in the heat transfer recovered in the condenser was calculated using equations (15) and (16), where the uncertainty associated with the temperature reading is estimated to be ±0.05% of the reading + 0.3°C and that in the flowmeter reading 2%:

SQ̇c=Q̇c√( SFRw

F Rw )2

+( SΔ T c

ΔT c )2

(9)

SΔ Tc=√ ( ST c ,o )2+(STc, i )

2 (10)

The uncertainty in the rate of heat transfer is presented in Figure 38 and Figure 39. The lowest uncertainty occurs at the highest inlet temperature of 250°C and the highest flow rate of 0.14 kg/s, most specifically in case 3. Considering equations (9) and (10), the smaller the temperature difference across the condenser, the higher the uncertainty. The figures also show that the uncertainty of the heat transfer in case 1 is the highest, when considering the lowest temperature difference across the condenser.

10. Conclusion

The investigation reported in this paper was successfully carried out using CFD as a numerical tool, validated both experimentally and theoretically. The validation was carried out on a different number of shell passes on the evaporator section for various air inlet conditions. The resistance analogy approach was applied to model the whole TSHX system in order to calculate the overall resistance. This approach was essential to replace the two-phase heat transfer inside the thermosyphons by pure conduction in ANSYS Fluent. The thermal conductivity was used as a boundary condition for the solid rods in Fluent, which successfully simulated the effect of the boiling and condensation heat transfer mechanisms. Also, the Realizable k−ε turbulence model was shown to be reliable in simulating the fluid behaviour and thermal performance of this type of heat exchanger system. For one shell pass, nucleate pool boiling was the only phenomenon taking place inside each thermosyphon water pool as the entire evaporator was filled with the working fluid. Adding more passes has increased the length of the evaporator section, which added one mode of heat transfer in the second and third passes, namely nucleate film boiling. Increasing the number of shell passes added more heat transfer modes in a parallel configuration and resulted in a decrease of the overall resistance and, therefore, an increase of the heat duty of the heat exchanger. The optimum performance of the TSHX was observed at the highest mass flow rate and inlet temperature (0.14kg/s, 250°C) for each case, 4403, 6191 and 9375 Watts for case 1, case 2 and case 3 respectively. The validated CFD model could be used as a tool to investigate shell side problems, including fouling in various heat exchanger designs as it provides a clear visualisation of the recirculation zones.

Acknowledgments

The reported research was funded by Innovate UK – Project number: 102716.

12

Nomenclature

A (m2) Heat transfer areaC (W/K) Heat capacity rateCr (-) Ratio of heat capacitiescp (J/(kg.K)) Specific heat capacityCsf (-) Coefficient of liquid/surface combinationh (W/(m2.K) Heat transfer coefficienthfg (J/kg) Specific enthalpy of evaporation / latent heat of vaporisationKp (-) Dimensionless parameterk (W/(m.K)) Thermal conductivityl (m) Lengthṁ (kg/s) Mass flow rateNu (-) Nusselt numberPr (-) Prandtl numberQ̇ (W) Heat transfer rateq̇ (W) Local heat transfer rateR (K/W) Thermal resistancerout (m) Thermosyphon outer radiusrin (m) Thermosyphon inner radiusRe (-) Reynolds numberT (oC) TemperatureT (oC) Average temperatureΔT (oC) Temperature differenceU (W/(m2.K)) Overall heat transfer coefficientε (-) Effectivenessμ (Pa.s) Dynamic viscosityρ (kg/m3) Densityσ (N/m) Surface tension

Subscriptsb Boilingc Condenser sideh Convectione Evaporator sidei Inletk ConductionL Liquid phaseLM Logarithmic meann Number of pipeso Outletr RateT TotalV Vapour phasew Water6TPs 6 thermosyphonsAbbreviations

TS ThermosyphonHP Heat PipeHPHX Heat Pipe based Heat ExchangerCFD Computational Fluid Dynamicsk-ε k-epsilon turbulence model

13

NTU Number of Transfer UnitsTPCT Two-Phase Closed Thermosyphon

References

[1] A. Al Makky, A. Alaswad, D. Gibson, A.G. Olabi, Prediction of the gas emission from porous media with the concern of energy and environment, Renew. Sustain. Energy Rev. 68 (2016) 1–13. doi:10.1016/j.rser.2016.08.001.

[2] Climate Change Act 2008, (2008). http://www.legislation.gov.uk/ukpga/2008/27/contents.

[3] European Environment Agency, (2016). http://www.eea.europa.eu/themes/climate/intro.

[4] A.G. Olabi, Developments in sustainable energy and environmental protection, Energy. 39 (2012) 2–5. doi:10.1016/j.energy.2011.12.037.

[5] J. Danielewicz, M.A. Sayegh, B. Śniechowska, M. Szulgowska-Zgrzywa, H. Jouhara, Experimental and analytical performance investigation of air to air two phase closed thermosyphon based heat exchangers, Energy. 77 (2014) 82–87. doi:10.1016/j.energy.2014.04.107.

[6] H. Shabgard, M.J. Allen, N. Sharifi, S.P. Benn, A. Faghri, T.L. Bergman, Heat pipe heat exchangers and heat sinks: Opportunities, challenges, applications, analysis, and state of the art, Int. J. Heat Mass Transf. 89 (2015) 138–158. doi:10.1016/j.ijheatmasstransfer.2015.05.020.

[7] A.G. Olabi, The 3rd international conference on sustainable energy and environmental protection SEEP 2009 – Guest Editor’s Introduction, Energy. 35 (2010) 4508–4509. doi:10.1016/j.energy.2010.09.053.

[8] A.G. Olabi, Hydrogen and Fuel Cell developments: An introduction to the special issue on “The 8th International Conference on Sustainable Energy and Environmental Protection (SEEP 2015), 11–14 August 2015, Paisley, Scotland, UK,” Int. J. Hydrogen Energy. 41 (2016) 16323–16329. doi:10.1016/j.ijhydene.2016.07.235.

[9] H. Jouhara, A.J. Robinson, Experimental investigation of small diameter two-phase closed thermosyphons charged with water, FC-84, FC-77 and FC-3283, Appl. Therm. Eng. 30 (2010) 201–211. doi:10.1016/j.applthermaleng.2009.08.007.

[10] H. Jouhara, T.K. Nannou, L. Anguilano, H. Ghazal, N. Spencer, Heat pipe based municipal waste treatment unit for home energy recovery, Energy. (2017). doi:10.1016/j.energy.2017.02.044.

[11] K. Negishi, T. Sawada, Heat transfer performance of an inclined two-phase closed thermosyphon, Int. J. Heat Mass Transf. 26 (1983) 1207–1213. doi:10.1016/S0017-9310(83)80175-6.

[12] H. Shabgard, B. Xiao, A. Faghri, R. Gupta, W. Weissman, Thermal characteristics of a closed thermosyphon under various filling conditions, Int. J. Heat Mass Transf. 70 (2014) 91–102. doi:10.1016/j.ijheatmasstransfer.2013.10.053.

[13] S. H. Noie, M. R. Sarmasti Emami, M. Khoshnoodi, Effect of inclination angle and filling ratio on thermal performance of an inclined two-phase closed thermosyphon under normal operating conditions, Heat Transf. Eng. 4 (2007) 365–371. doi:10.1080/01457630601122997.

[14] H. Jouhara, Z. Ajji, Y. Koudsi, H. Ezzuddin, N. Mousa, Experimental investigation of an inclined-condenser wickless heat pipe charged with water and an ethanol–water azeotropic mixture, Energy. 61 (2013) 139–147. doi:10.1016/j.energy.2012.09.033.

[15] H. Jouhara, H. Merchant, Experimental investigation of a thermosyphon based heat exchanger used in energy efficient air handling units, Energy. 39 (2012) 82–89. doi:10.1016/j.energy.2011.08.054.

14

[16] S.H. Noie-Baghban, G.R. Majideian, Waste heat recovery using heat pipe heat exchanger (HPHE) for surgery rooms in hospitals, Appl. Therm. Eng. 20 (2000) 1271–1282. doi:10.1016/S1359-4311(99)00092-7.

[17] J. Ramos, A. Chong, H. Jouhara, Experimental and numerical investigation of a cross flow air-to-water heat pipe-based heat exchanger used in waste heat recovery, Int. J. Heat Mass Transf. 102 (2016) 1267–1281. doi:10.1016/j.ijheatmasstransfer.2016.06.100.

[18] H. Mroue, J.B. Ramos, L.C. Wrobel, H. Jouhara, Experimental and numerical investigation of an air-to-water heat pipe-based heat exchanger, Appl. Therm. Eng. 78 (2015) 339–350. doi:10.1016/j.applthermaleng.2015.01.005.

[19] B. Fadhl, L.C. Wrobel, H. Jouhara, Numerical modelling of the temperature distribution in a two-phase closed thermosyphon, Appl. Therm. Eng. 60 (2013) 122–131. doi:10.1016/j.applthermaleng.2013.06.044.

[20] H. Jouhara, B. Fadhl, L.C. Wrobel, Three-dimensional CFD simulation of geyser boiling in a two-phase closed thermosyphon, Int. J. Hydrogen Energy. 41 (2016) 16463–16476. doi:10.1016/j.ijhydene.2016.02.038.

[21] B. Fadhl, L.C. Wrobel, H. Jouhara, CFD modelling of a two-phase closed thermosyphon charged with R134a and R404a, Appl. Therm. Eng. 78 (2015) 482–490. doi:10.1016/j.applthermaleng.2014.12.062.

[22] Y. You, A. Fan, S. Huang, W. Liu, Numerical modeling and experimental validation of heat transfer and flow resistance on the shell side of a shell-and-tube heat exchanger with flower baffles, Int. J. Heat Mass Transf. 55 (2012) 7561–7569. doi:10.1016/j.ijheatmasstransfer.2012.07.058.

[23] D.A. Reay, P.A. Kew, R.J. McGlen, Chapter 3 – Heat pipe components and materials, in: Heat Pipes, 2014: pp. 65–94. doi:10.1016/B978-0-08-098266-3.00003-0.

[24] J. Danielewicz, B. Śniechowska, M.A. Sayegh, N. Fidorów, H. Jouhara, Three-dimensional numerical model of heat losses from district heating network pre-insulated pipes buried in the ground, Energy. 108 (2016) 172–184. doi:10.1016/j.energy.2015.07.012.

[25] L.M. Poplaski, A. Faghri, T.L. Bergman, Analysis of internal and external thermal resistances of heat pipes including fins using a three-dimensional numerical simulation, Int. J. Heat Mass Transf. 102 (2016) 455–469. doi:10.1016/j.ijheatmasstransfer.2016.05.116.

[26] Lecture 6 Turbulence Modeling, Introduction to ANSYS FLUENT, ANSYS, Inc., Release 13.0, (2010). http://imechanica.org/files/fluent_13.0_lecture06-turbulence.pdf.

[27] Lecture 7: Turbulence Modeling ANSYS Inc 16.0 Release, (2015). https://www.scribd.com/document/302199566/Fluent-Intro-16-0-L07-Turbulence.

[28] F.P. Incropera, D.P. DeWitt, Introduction to Heat Transfer, Wiley, 2011. https://books.google.co.uk/books?id=kvawuAAACAAJ.

15

Figure 1 Thermosyphon schematic.

16

Figure 2 Test unit dimensions.

17

Figure 3 Top view of the condenser (all dimensions are in mm).

18

Figure 4 One, two and three air passes (left to right).

19

Figure 5 Air (evaporator) and water (condenser) circuit.

20

Figure 6 Thermocouple locations (Each red dot represents a K-type thermocouple).

21

Figure 7 Resistance network, Case 3.

22

Table 1 Thermal resistance correlations.

Resistance Heat transfer mode Analytical/empirical

Rh , Air−Wall Air-Wall convection1

hconvection A e

(11)

Rk , WallConduction across thermosyphon wall

ln( rout

r ¿)

2 πkL

(12)

Rh , NPB Nucleate pool boiling q̇A

=μ hfg √ g(ρL−ρv )σ ( c p ∆ T b

h fg C sf )1m Pr

−nm (

13)

Rh , NFB Nucleate film boilingq̇A

=(1.155 ×10−3 ) N uμf

0.33 Pr0.35 K p0.7( qe lm

ρV h fg v l)

0.7 ∆ Tb kll

(14)

Rh , NFC Film condensation hm=0.943 ×( k l3 ρL ( ρL−ρV ) g hfg

μl θl )1 /4

(15)

Rh , Water−WallWater-Wall convection

1hconvection A c

(16)

23

Figure 8 Resistance network, Case 1 (left) and Case 2 (right).

24

Table 2 Mesh parameters for case 3.

Coarse Medium FineRelevance 0 50 100Nodes 650,472 1,600,751 2,328,476Elements 2,147,026 4,942,635 8,009,957Maximum skewness 0.88 0.89 0.84Average skewness 0.228 0.220 0.211

25

Figure 9 Medium mesh of the whole system (case 3).

26

Figure 10 Top view of the condenser: Coarse mesh (left) and fine mesh (right).

27

0.02 0.05 0.08 0.11 0.14 0.17-0.5%

0.5%

1.5%

2.5%

3.5%

4.5%

Coarse

Medium

Fine

Mass flow rate (kg/s)

% E

rror

Figure 11 % error at 150°C for the air outlet temperatures (case 3).

0.02 0.05 0.08 0.11 0.14 0.17

-12%

-10%

-8%

-6%

-4%

-2%

0%

2%

4%

Coarse

Medium

Fine

Mass flow rate (kg/s)

% E

rror

Figure 12 % error at 150°C for the water outlet temperatures (case 3).

0.02 0.05 0.08 0.11 0.140%

1%

2%

3%

4%

5%

6%

7%

8%

9%

10%

CoarseMediumFine

Mass flow rate (kg/s)

% E

rror

Figure 13 % error at 250°C for the air outlet temperatures (case 3).

0.02 0.05 0.08 0.11 0.14

-20%

-15%

-10%

-5%

0%

Coarse

Medium

Fine

Mass flow rate (kg/s)

% E

rror

Figure 14 % error at 250°C for the water outlet temperatures (case 3).

28

Table 3 Boundary conditions.

BC Type Mass Flow Inlet (kg/s) Inlet Temperature (°C)Air Inlet Mass flow inlet 0.05 to 0.14 at

increments of 0.03100 to 250 at increments

of 50Air Outlet Pressure outlet — —Water Inlet Mass flow inlet 0.08 ≈15Water Outlet Pressure outlet — —

29

0.05 0.08 0.11 0.140

1000

2000

3000

4000

5000

6000

7000

8000

100%

1st pass

Air mass flow rate (kg/s)

Heat

Tran

sfer R

ate (W

)

Figure 15 Case 1 at 150°C.

0.05 0.08 0.11 0.140

1000

2000

3000

4000

5000

6000

7000

8000

100%

1st pass

Air mass flow rate (kg/s)

Heat

Tran

sfer R

ate (W

)

Figure 16 Case 1 at 250°C.

0.05 0.08 0.11 0.140

1000

2000

3000

4000

5000

6000

7000

8000

63% 63% 64% 65%37%

37% 36% 35%

2nd pass 1st pass

Air mass flow rate (kg/s)

Heat

Tran

sfer R

ate (W

)

Figure 17 Case 2 at 150°C.

0.05 0.08 0.11 0.140

1000

2000

3000

4000

5000

6000

7000

8000

65%64% 64% 64%

35%

36%36%

36%

2nd pass 1st pass

Air mass flow rate (kg/s)

Heat

Tran

sfer R

ate (W

)

Figure 18 Case 2 at 250°C.

0.05 0.08 0.11 0.140

1000

2000

3000

4000

5000

6000

7000

8000

49% 52% 53% 54%30%

29%28% 27%

21%

20%19% 18%

3rd pass 2nd pass

1st pass

Air mass flow rate (kg/s)

Heat

Tran

sfer R

ate (W

)

Figure 19 Case 3 at 150°C.

0.05 0.08 0.11 0.140

1000

2000

3000

4000

5000

6000

7000

8000

52%53%

55%57%

29%

30%

29%29%

19%

17%

16%

14%3rd pass 2nd pass

1st pass

Air mass flow rate (kg/s)

Heat

Tran

sfer R

ate (W

)

Figure 20 Case 3 at 250°C.

30

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 100000

1000

2000

3000

4000

5000

6000

7000

8000

9000

10000

Theoretical CFD

Experimental +15%

-15%

Experimental heat transfer rate (W)

Theo

retica

l / C

FD h

eat tr

ansfe

r rate

(W)

Figure 21 Comparison between experimental, theoretical and CFD results (case 1).

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 100000

1000

2000

3000

4000

5000

6000

7000

8000

9000

10000

Theoretical CFD

Experimental +15%

-15%

Experimental heat transfer rate (W)

Theo

retica

l / CF

D he

at tra

nsfer

rate

(W)

Figure 22 Comparison between experimental, theoretical and CFD results (case 2).

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 100000

1000

2000

3000

4000

5000

6000

7000

8000

9000

10000

Theoretical CFD

Experimental +15%

-15%

Experimental heat transfer rate (W)

Theo

retica

l / CF

D he

at tra

nsfer

rate

(W)

Figure 23 Comparison between experimental, theoretical and CFD results (case 3).

0.02 0.05 0.08 0.11 0.14 0.170%

5%

10%

15%

20%

25%

Case 1 Case 2

Case 3

Mass flow rate (kg/s)

% Er

ror

Figure 24 Percentage error of the evaporator outlet temperature at different air mass flow rates.

0.02 0.05 0.08 0.11 0.14 0.170%

5%

10%

15%

20%

25%

Case 1 Case 2

Case 3

Mass flow rate (kg/s)

% E

rror

Figure 25 Percentage error of the condenser outlet temperature at different air mass flow rates.

31

50 70 90 110 130 150 170 190 210 230 25050

70

90

110

130

150

170

190

210

230

250

0.05 kg/s

0.08 kg/s

0.11 kg/s

0.14 kg/s

y = x

Evaporator outlet T in °C (Experimental)

Evap

orato

r ou

tlet T

in °C

(CFD

)

Figure 26 Evaporator outlet temperature for CFD and experimental at different inlet conditions (case 1).

0 5 10 15 20 25 30 35 40 45 500

5

10

15

20

25

30

35

40

45

50

0.05 kg/s

0.08 kg/s

0.11 kg/s

0.14 kg/s

y = x

Evaporator outlet T in °C (Experimental)

Cond

ense

r outl

et T

in °C

(CFD

)

Figure 27 Condenser outlet temperature for CFD and experimental at different inlet conditions (case 1).

50 70 90 110 130 150 170 190 210 230 25050

70

90

110

130

150

170

190

210

230

250

0.05 kg/s

0.08 kg/s

0.11 kg/s

0.14 kg/s

y = x

Evaporator outlet T in °C (Experimental)

Evap

orato

r out

let T

in °C

(CFD

)

Figure 28 Evaporator outlet temperature for CFD and experimental at different inlet conditions (case 2).

0 5 10 15 20 25 30 35 40 45 500

5

10

15

20

25

30

35

40

45

50

0.05 kg/s

0.08 kg/s

0.11 kg/s

0.14 kg/s

y = x

Condenser outlet T in °C (Experimental)

Cond

ense

r outl

et T

in °C

(CFD

)

Figure 29 Condenser outlet temperature for CFD and experimental at different inlet conditions (case 2).

50 70 90 110 130 150 170 190 210 230 25050

70

90

110

130

150

170

190

210

230

250

0.05 kg/s

0.08 kg/s

0.11 kg/s

0.14 kg/s

y = x

Evaporator outlet T in °C (Experimental)

Evap

orato

r outl

et T

in °C

(CFD

)

Figure 30 Evaporator outlet temperature for CFD and experimental at different inlet conditions (case 3).

0 5 10 15 20 25 30 35 40 45 500

5

10

15

20

25

30

35

40

45

50

0.05 kg/s

0.08 kg/s

0.11 kg/s

0.14 kg/s

y = x

Condenser outlet T in °C (Experimental)

Cond

ense

r ou

tlet T

in °C

(CFD

)

Figure 31 Condenser outlet temperature for CFD and experimental at different inlet conditions (case 3).

32

Figure 32 Temperature contours across the evaporator for the three cases at 150°C, 0.14 kg/s.

33

Figure 33 Velocity vector plot across the evaporator for the three cases at (150°C, 0.14 kg/s).

34

Figure 34 Top view of condenser temperature contour for the three cases at (150°C, 0.14kg/s) and z=1.9m.

35

0.02 0.05 0.08 0.11 0.14 0.1710%

15%

20%

25%

30%

35%

40%

250 °C

200 °C

150 °C

100 °C

Air mass flow rate (kg/s)

Effe

ctive

ness

Figure 35 Effectiveness at different air inlet temperature and mass flow rates (Case 3).

0.02 0.07 0.12 0.17 0.22 0.27 0.32 0.37 0.42 0.470

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

Case 1 Case 2

Case 3

NTU

Effe

ctive

ness

Figure 36 Effectiveness vs NTU at different inlet conditions (100°C).

0.02 0.07 0.12 0.17 0.22 0.27 0.32 0.37 0.42 0.470

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

Case 1 Case 2

Case 3

NTU

Effe

ctive

ness

Figure 37 Effectiveness vs NTU at different inlet conditions (250°C).

0.02 0.05 0.08 0.11 0.14 0.170%

5%

10%

15%

20%

25%

Case 1 Case 2

Case 3

Air mass flow rate (kg/s)

Unce

rtain

ty

Figure 38 Uncertainty of the three cases at 100°C air inlet temperature.

0.02 0.05 0.08 0.11 0.14 0.170%

5%

10%

15%

20%

25%

Case 1 Case 2

Case 3

Air mass flow rate (kg/s)

Unce

rtain

ty

Figure 39 Uncertainty of the three cases at 250°C air inlet temperature.

36