3.1. Evaporator
The evaporator, a crucial component in the gravity loop heat pipe (GLHP) heat transfer process, plays a pivotal role in absorbing heat from the source and inducing the phase change of the working fluid. The heat transfer mechanisms within the evaporator are multifaceted.
Firstly, direct contact with the heat source occurs through a conductive interface, enabling the conduction of heat from the source to the evaporator wall. The resulting temperature gradient drives the heat transfer process. Secondly, the heat conducted through the evaporator wall is convectively transferred to the working fluid, and the design of the evaporator wall maximizes the contact area with the working fluid, optimizing heat transfer efficiency. As the working fluid absorbs heat, its temperature rises until it reaches the saturation temperature corresponding to the operating pressure of the evaporator. At this point, the liquid working fluid undergoes a phase change, vaporizing and transforming into steam, a process accompanied by the absorption of heat known as latent heat of vaporization. This phase change phenomenon facilitates efficient heat transfer, allowing significant amounts of heat to be transferred with relatively small temperature differences.
In summary, the heat transfer in a GLHP evaporator is a synergistic interplay of conduction, convection, and phase change mechanisms, enabling the efficient absorption of heat from the source and the conversion of the liquid working fluid into vapor, thereby facilitating efficient cooling and thermal management.
The flow of flue gases in the evaporator follows a specific path, initially along the outer wall and then along the inner wall. This longitudinal bypass of the evaporator facilitates heat transfer through convection between the flue gases and the inner and outer surfaces of the evaporator shell. The evaporator walls are characterized by their smooth and straight configuration, allowing for the consideration of heat transfer through a flat wall when analyzing the convective heat transfer on the outer shell. On the inner side of the evaporator, heat transfer in the annular region needs to be accounted for due to the presence of a deflector. Furthermore, the evaporator receives heat from the fireplace insert through radiation. Additionally, within the evaporator, the condensate undergoes boiling, which further contributes to the heat transfer process.
Flue gases heat the outer wall of the evaporator, flow through the inner ring, heat the inner wall, and continue to the chimney. The total heat transfer in this system is calculated using Equation (1).
where
Qc is the total heat transfer by convection, and
She is the surface area of the heat exchanger. Convection facilitates efficient heat transfer from the flue gases to the evaporator walls. The heat transfer rate
q is calculated using Fourier’s law of thermal conduction (2).
The heat transfer coefficient ‘
k’ is determined by the Equation (3), where
λ represents thermal conductivity, and
dss represents stainless-steel conductivity. This coefficient characterizes the efficiency of heat transfer and is crucial for analyzing thermal conductivity in a system.
The wall of the evaporator is exposed to flue gases, which in turn heat the wall. This transferred heat is then efficiently utilized to raise the temperature of the working medium (
Figure 3). The process involves the conduction of heat through the wall, enabling the efficient exchange of thermal energy from the flue gases to the working medium, ultimately contributing to the overall performance and functionality of the system.
The heat transfer coefficient ‘
k’ can be calculated from the Nusselt number using Equation (4).
The calculation of thermal conductivity
λ involves considering the thermokinetic temperature of the boundary layer
Tδ (5). This temperature plays a significant role in determining the thermal conductivity value.
The average surface loss coefficient ε for a given length
L can be determined using Equation (6), where
ReL denotes the Reynolds number based on the length
L. This coefficient characterizes the average surface losses within the flow and is derived from scientific principles. It provides valuable insights into the heat transfer behavior in the system.
The Reynolds number, denoted as
ReL, is a dimensionless parameter used to characterize the flow regime. For laminar flow conditions (7), w represents the velocity of the fluid,
L is the characteristic length, and
v denotes the kinematic viscosity of the fluid. This criterion indicates that the flow is considered laminar when the Reynolds number falls below the specified threshold value of 3 × 10
5.
The Prandtl number
Pr is a dimensionless quantity that characterizes the relative importance of momentum and heat transfer in a fluid. It is defined as the ratio of kinematic viscosity
ν to thermal diffusivity
α (8).
For flow conditions where the Reynolds number
ReL is less than 10
5 and the Prandtl number
Pr falls within the range of 0.1 to 1000, the Nusselt number
Nu can be determined using the correlation Equation (9).
This equation relates the Nusselt number to the Reynolds number and Prandtl number, capturing the combined effects of convective heat transfer and fluid properties. It provides an estimation of the convective heat transfer coefficient based on these dimensionless parameters.
The Nusselt number
Nu for flow on the inner wall of the evaporator, with a Reynolds number
Re below 2300, is determined by Equation (10). To ensure accurate application of this equation, it is important to have a Grashof number greater than 25,000 and a product of
Re,
Pr, and (
D/
L) equal to or greater than 7.17. The viscosity ratio (
η/
ηs) accounts for the ratio of fluid viscosities between the working fluid and the reference fluid.
Equation (11) provides valuable insights into the convective heat transfer behavior between a solid surface and a fluid medium. It considers factors such as the thermal conductivity
λ, the Prandtl number
Pr, and the flow characteristics including velocity
w and characteristic length
L. This equation allows for a better understanding of the average convective heat transfer coefficient
α over a given length
L.
The transfer of heat from the fireplace insert to the evaporator in a closed system involves radiation. The heat transfer is influenced by the surfaces of the evaporator
A1 and the small heat source
A2. The external radiation heat transfer, denoted as
Qr,ext, is determined by the emissivity of the surfaces ε, which affects their ability to emit and absorb thermal radiation. The equation for
Qr,ext is given by Equation (12), where the temperatures of the evaporator and the fireplace insert are denoted as
T1 and
T2, respectively, and
Co is a constant.
The overall emissivity, ε
12, is determined by the individual emissivities of the surfaces ε
1 and ε
2 and their respective areas
A1 and
A2. Radiant flux between surfaces
A1 and
A2 forming a closed system is shown on
Figure 4.
The relationship for ε
12 (13) considers the interaction of the two surfaces in the radiation heat transfer process.
Heat transfer through boiling is a common phenomenon characterized by the formation and growth of vapor bubbles on a heated surface [
28,
29]. The heat transfer in this process is influenced by factors such as the Reynolds number
ReB, Prandtl number
Prk, and pressure
p. In the case of bubble boiling, there are two criterial equations that describe the Nusselt number
NuB for different ranges of Reynolds number. Equation (14) applies to
,while Equation (15) applies to
.
These equations provide correlations between NuB, ReB, and Prk, allowing for the estimation of heat transfer performance in boiling systems within the specified range. It is important to note that the Prandtl number should be between 0.86 and 7.6, and the pressure should range from 4500 to 17.5 × 106 Pa. These criterial equations serve as valuable tools in the design and optimization of heat transfer systems involving boiling, enhancing our understanding and control of heat transfer processes in such applications.
In boiling processes, the Nusselt number
NuB is a key parameter that quantifies the convective heat transfer. It is defined by Equation (16), where h represents the convective heat transfer coefficient, B is a characteristic length, and
λk is the thermal conductivity of the fluid.
The Reynolds number
ReB characterizes the fluid flow behavior in boiling. It is determined using Equation (17), where
q is the heat flux,
B is a characteristic length,
lb is the latent heat,
ρs is the density of the fluid, and
νc is the kinematic viscosity.
The Prandtl number
Prk describes the relative importance of momentum diffusion to thermal diffusion in boiling. It is determined by Equation (18), where
νc is the kinematic viscosity and
αc is the thermal diffusivity of the fluid. The Prandtl number helps in understanding the relative rate of momentum and heat transfer during boiling.
The variable
B in the equations represents a characteristic length scale, which is used to normalize the effects of size or geometry in the boiling process. It is defined by Equation (19), where
cs is the specific heat capacity,
ρc is the density of the solid,
σ is the Stefan–Boltzmann constant,
Ts is the surface temperature,
lb is the latent heat, and
ρc is the density of the fluid. The characteristic length
B accounts for the specific thermodynamic properties of the system and enables a better understanding of the heat transfer phenomena during boiling processes.
The determination of the heat transfer coefficient in bubble boiling is influenced by the pressure conditions. For pressures up to
p = 10
5 Pa, the heat transfer coefficient
h can be calculated using Equations (20) and (21), where ∆
T is the temperature difference and
q is the heat flux.
For pressures ranging from 0.01 to 15 MPa, the heat transfer coefficient is given by Equations (22) and (23). These equations provide a means to estimate the heat transfer coefficient on the basis of the pressure and temperature difference, allowing for analysis and design of bubble boiling systems.
3.2. Condenser
Heat transfer in the GLHP condenser plays a vital role in efficient thermal management by facilitating the release of heat from steam to combustion air. This process involves several key mechanisms that contribute to heat transfer within the condenser.
Firstly, hot vapor from the evaporator enters the condenser and contacts the condenser wall, which is at a lower temperature. Heat is then conducted from the steam to the condenser wall. The temperature difference between the vapor and the condenser wall drives this heat transfer process.
Secondly, as the vapor encounters the cooler condenser wall, it undergoes a phase change from vapor to liquid through condensation. This phase change releases latent heat of condensation, which is the thermal energy associated with the transition. The condenser wall effectively absorbs this latent heat, contributing to the overall heat transfer process. The condensed liquid resulting from condensation flows down the condenser wall under the influence of gravity, returning to the evaporator section. This flow ensures a continuous supply of liquid working fluid to the evaporator, thereby sustaining the operation of the GLHP system.
Overall, heat transfer in the GLHP condenser involves conduction, condensation, and gravity-driven liquid flow mechanisms. These processes work in tandem to extract heat from the vapor and release it to the combustion air, enabling effective cooling and thermal regulation.
The condenser of the GLHP system is cooled by the supplied combustion air, with heat transfer occurring as the air flows transversely around the condenser. The condenser is typically constructed using a ribbed tube, comprising both a smooth section and ribs. Heat conduction takes place through the folded cylindrical wall of the condenser, allowing for efficient transfer of thermal energy. Functioning as a four-stroke heat exchanger with one-sided mixed flow, the condenser ensures effective heat transfer within the GLHP system.
The condenser comprises a linear array of four ribbed tubes, and the heat transfer efficiency and pressure drop in the tube bundle with external fins are influenced by a multitude of geometric parameters. To assess the hydraulic and thermal performance of the condenser, mathematical relationships that incorporate the geometric characteristics of the individual tubes and the tube bundle can be employed.
One such correlation is the Nusselt number in Equation (24). This equation encapsulates the heat transfer coefficient
Nu, the mass velocity on the tube side
Gα, the Reynolds number
Re, and the Prandtl number
Pr.
Furthermore, relation (25) provides insights into the heat transfer surface area (
α ×
S) and the size of the tube bundle
C. Here,
S1,
nr,
l1,
ϕ,
Sr,
St, and
ηr represent geometric parameters associated with the arrangement of the tube bundle.
The coefficients
Kα and
n are determined on the basis of the specific geometric attributes of the tubes (
Table 1). For the specific configuration of tubes arranged in a linear fashion, the value of
Gα is unity. This simplification is applicable to the arrangement of the four ribbed tubes within the condenser. By considering these geometric parameters and employing the pertinent equations, the hydraulic and thermal characteristics of the condenser system can be effectively evaluated with enhanced precision.
In the context of fluid dynamics, the Reynolds number
Re is determined at the narrowest section of the tube bundle. It is defined as the ratio of the velocity
w to the kinematic viscosity
v (26). Here,
de represents the equivalent diameter of the ribbing, which is considered a characteristic dimension.
The hydraulic diameter
dh is employed to estimate
de and is calculated using relationship (27). In this expression,
St denotes the total surface area,
dt refers to the tube diameter,
Sr represents the rib surface area, and
nr corresponds to the number of ribs per tube.
The Nusselt number
Nu characterizing the heat transfer in this situation can be expressed as (28), where
C represents a constant,
ρv is the density of the vapor,
g denotes the acceleration due to gravity,
ρp corresponds to the density of the fluid,
hfg represents the average latent heat of vaporization,
D represents the characteristic length or diameter,
μv is the dynamic viscosity of the vapor,
kv denotes the thermal conductivity of the vapor,
Tsat is the saturation temperature, and
Ts represents the surface temperature.
The addition of aluminum rolled ribs to a smooth copper tube enhances heat transfer by increasing the surface area available for heat exchange. This improvement is achieved through a combination of conduction and convection mechanisms.
Conduction enables efficient heat transfer from the inner surface of the copper tube to the outer surface, allowing heat to flow into the aluminum ribs. The high thermal conductivity of copper facilitates this process, promoting effective conduction across the interface.
Convective heat transfer occurs as fluid flows over the outer surface of the ribbed structure. The presence of the aluminum ribs disrupts the fluid flow, creating turbulent boundary layers that enhance convective heat transfer. This turbulent flow promotes efficient heat exchange between the fluid and the aluminum ribs.
The combination of conduction and convective heat transfer between the copper tube and the aluminum ribs results in improved heat transfer performance. By increasing the available surface area, the addition of aluminum rolled ribs enhances the overall efficiency of heat exchangers.
The cylindrical wall consists of two layers, and the equations describe the heat transfer across the wall. Equation (29) represents heat transfer from the inner surface to the outer surface, while (30) represents heat transfer from the outer surface to the surroundings. These equations determine temperature differences and heat flow rates.
To calculate the temperature difference across the entire wall, Equation (31) is used. This accounts for heat flow rates and thermal conductivities of both layers, providing insight into temperature distribution.
The equivalent thermal conductivity
λₑₖᵥ of the composite wall is determined by (32). This equation provides the overall thermal conductivity of the composite wall, considering individual conductivities and geometries of the layers.
Equations (30)–(32) enable analysis of heat transfer in cylindrical walls with multiple layers, providing valuable insights into temperature distribution, heat flow rates, and effective thermal conductivity. The dimensions appearing in relations (30)–(32) are illustrated in
Figure 5.
The heat transfer process in the condenser involves the direct flow of combustion air perpendicular to the axis of the finned tubes. This airflow configuration promotes efficient heat exchange by allowing the air to mix with the surrounding environment. However, the walls of the condenser tubes act as barriers, preventing the partial streams of steam from mutually mixing within the condenser.
As a result, the condenser operates as a four-stroke heat exchanger with a one-sided mixed crossflow. The combustion air flows across the condenser tubes, absorbing heat from the steam as it passes through the finned surfaces. This one-sided mixed crossflow design ensures effective heat transfer between the steam and the surrounding air, facilitating the release of thermal energy from the condenser.
In heat exchanger analysis, the logarithmic temperature difference method is employed to determine temperature distributions and temperature differences across the exchanger. The method utilizes the concept of a logarithmic temperature difference to obtain more accurate estimations of heat transfer performance.
The temperature at the first fluid outlet,
ts1, can be calculated using Equation (33)
Similarly, the temperature at the second fluid outlet,
ts2, can be determined by Equation (34).
The mean temperature difference, Δ
tm, can be determined by Equation (35).
In Equations (33)–(35), t1 and t2 represent the inlet temperatures of the first and second fluids, α1 and α2 are the heat transfer coefficients of the respective fluids, S1 and S2 are the heat transfer surface areas on the respective sides, D1 and D2 are the diameters of the channels, L is the length of the heat exchanger, and λekv represents the equivalent thermal conductivity of the composite wall.
The mean logarithmic slope method provides a more accurate evaluation of temperature differences and facilitates the analysis and design of heat exchangers. By considering the logarithmic average of temperature differences, it accounts for variations in temperature distribution across the exchanger, enabling improved heat transfer calculations.
3.3. Results
The heat output of the evaporator is influenced by the rising flue gas temperature (
Figure 6). As the flue gas temperature increases, all heat outputs exhibit a diminishing trend. Amongst these outputs, the dominant contribution comes from the radiative heat transfer rate,
Qr,int, which acts upon the inner wall of the evaporator. Remarkably, this radiative heat transfer rate surpasses the convective heat outputs,
Qk,ext and
Qk,int, on the outer and inner surfaces of the evaporator, respectively. Furthermore, the heat outputs on the inner surface of the evaporator demonstrate higher magnitudes compared to those observed on the outer shell of the evaporator.
The logarithmic temperature difference method (LDMT) was validated by comparing it with the operational characteristic function (char) to evaluate heat transfer in the condenser. Both methods yielded consistent results, indicating a reliable estimation of heat transfer (
Figure 7).
With increasing steam temperature, the temperature of the combustion air t22 also rose. Consequently, the temperature difference between the combustion air and steam increased. This larger temperature difference enhanced the efficiency of heat transfer between the condenser and the combustion air, leading to improved heat exchange performance.
The mass flow rate of the working fluid varied with the change in flue gas temperature in both the evaporator and the condenser (
Figure 8). By overlaying the curves of the condensate mass flow rate ṁ
c and the steam mass flow rate ṁ
s, a point was identified with equal temperature and pressure, where the amount of evaporated condensate was equal to the amount of condensed condensate. The cooperation between the evaporator and the condenser is contingent upon maintaining identical temperature and pressure in both devices. The collaboration of the evaporator and the condenser occurs at a steam temperature of 160 °C and a saturated steam pressure of 618.3 kPa.
On the basis of the initial parameters used in the device’s design, we determined the efficiency of the small heat source. Subsequently, the mathematical model provided insights into the cooperative behavior of the evaporator and the condenser, allowing us to evaluate the device’s influence on enhancing the efficiency of the small heat source.
The mathematical model facilitated a recalibration of the small heat source’s efficiency, incorporating the parameters obtained from the simulation results. The model’s predictions demonstrated that the proposed device led to a significant improvement in the efficiency of the small heat source. The efficiency gains achieved were notable, with an increase of up to 3%. This enhancement is of great significance for small-scale heat sources, as even slight improvements can result in significant energy savings and operational cost reductions.
Moreover, the mathematical model’s comprehensive overview provided valuable insights into the underlying heat transfer mechanisms and interactions within the device. This deeper understanding of the system’s behavior enables further optimizations and potential modifications to achieve even higher efficiency gains.
In conclusion, our mathematical model demonstrated the considerable impact of the proposed device on improving the efficiency of small heat sources. This study serves as a foundation for future research and development in the field of thermal management and energy efficiency, offering promising possibilities for sustainable heat utilization and economic benefits.