5
Natural Convection Heat Transfer

5.1. Introduction

All of the situations studied in the previous chapters concern flow in forced convection, that is to say, the fluid motion is generated by an external unit: pump, compressor, etc. Yet, whilst such situations represent the majority of flows encountered in industrial systems, it is important to highlight the existence, in practice, of cases where motion is generated in the absence of any external element.

Indeed, unlike forced convection, natural (or free) convection corresponds to a heat transfer mode where the fluid motion is quite simply induced by forces of Archimedes, generated by differences in fluid density, the latter being themselves caused by temperature differences within the fluid. Examples of this type of flow can be found in the operation of flat thermosiphon solar collectors or in heat exchanges between wall-mounted heating radiators and ambient air.

5.2. Characterizing the motion of natural convection

Consider a fluid at temperature θ, in contact with a surface of temperature θw, greater than θ (see Figure 5.1). The fluid masses located in the vicinity of the surface heat up and their densities, thus, decrease.

This decrease in density makes these masses lighter which enables them to rise upward, with a force proportional to the density difference thus created. Indeed, the force per unit volume exerted on these masses is equal to math. As a result, a motion is generated, where the hot fluid masses rise, leaving space for the cooler masses, which “plunge” towards the hot plate. The acceleration, math, of this motion is obtained by dividing force by mass, or: math.

image

Figure 5.1. Natural convection. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

In order to derive an expression for the relative variation of density, math, let us introduce the constant-pressure thermal expansion coefficient:

math

Yet: math.

Or, as a first approximation: math.

Hence: math.

The acceleration, math, of the upward motion due to the temperature difference then becomes:

math

Thus, the motion will be conditioned by the parameter β(θw −θ)g, which constitutes the driving force for natural convection. From this parameter, we can then define a dimensionless number, called the Grashof number, which enables us to compare this driving force of natural convection (β(θw −θ)g) to the viscous forces determined by the kinematic viscosity math of the fluid considered.

The Grashof number is thus defined by: math, where:

δ is a characteristic dimension of the surface considered: the height, length, width or diameter, as appropriate;

ρ is the density of the fluid;

G is the acceleration of gravity;

β is the constant-pressure thermal expansion coefficient of the fluid:

math

μ is the viscosity of the fluid within the range of temperatures considered;

w −θ) is the gradient between the wall temperature, θw, and the fluid approach temperature, θ.

5.3. Correlations in natural convection

We have shown, through the application of dimensional analysis (see Chapter 1, section 1.4.2.2), that the natural convection heat transfer coefficient, h, i.e. the Nusselt number, can be expressed as a function of the Grashof and Prandtl numbers, as follows:

math

where:

math

Parameters C, a and b, are determined based on experimentation

Several experiments of this type have been conducted and have enabled appropriate parameters to be determined for the experimental situations considered. The following sections present the most significant correlations determined in this way.

The relations presented in the sub-sections below consider the different situations that can be of interest in engineering calculations:

  • – vertical plane surfaces;
  • – vertical cylindrical surfaces;
  • – horizontal surfaces;
  • – horizontal cylinders;
  • – plane surfaces forming an angle, α, with the horizontal surface;
  • – spheres;
  • – vertical conical surfaces;
  • – any vertical surface;
  • – volumes limited by parallel surfaces;
  • – cylinders, disks or spheres in rotation.

5.4. Vertical plates subject to natural convection

This type of problem corresponds to the scenario of wall-mounted ambient-heating radiators, where heat is transmitted from the radiator to the air in the room by natural convection. It is also encountered in the case of electronic-component heat sinks; see sections 5.17 to 5.19.

In industrial environments, several situations are encountered where vertical surfaces are cooled or heated by natural convection.

Both Gebhart (1961) and Kato, Nishiwaki and Hirata (1968) analyzed the problems relating to natural convection heat transfer between a significant mass of fluid at average temperature θm and hot vertical plane surfaces (temperature θpm > θm).

image

Figure 5.2. Vertical plate. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

Under these conditions, the following correlations have been determined:

  • – For math;
  • – For math.

In these relations:

  • – The characteristic dimension taken into account in the Nusselt and Grashof numbers is the height, L, of the surface:
    math
  • math;
  • – The physical properties of the liquid are taken at the average temperature of the film, defined by:
    math
    • - θPm is the average temperature of the plate;
    • - θm is the average temperature of the fluid.

5.5. Inclined plates subject to natural convection

image

Figure 5.3. Inclined plate. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

In order to estimate the natural convection heat transfer coefficient between a significant mass of fluid at average temperature θm and a flat plate forming an angle, α , with the horizontal surface, the following correlations can be used:

  • – for math;
  • – for math.

In these relations:

  • – the characteristic dimension taken into account in the Nusselt and Grashof numbers is the Length, L, of the surface:
    math
  • math;
  • the physical properties of the liquid are taken at the average temperature of the film, defined by: math:
    • - θPm is the average temperature of the plate;
    • - θm is the average temperature of the fluid.

5.6. Horizontal plates subject to natural convection

Several studies have been conducted on natural convection heat transfer between horizontal plates and the surrounding environment (Brown and Marco, 1958; Jakob, 1957). The correlations obtained are of great use in calculations relating to thermal design in buildings and particularly for the design of underfloor heating.

5.6.1. Case of underfloor heating

image

Figure 5.4. Underfloor heating. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

For a hot surface placed in the lower part of a chamber containing the fluid, one of the following correlations can be used:

  • – for math;
  • – for math.

In these relations, the characteristic dimension, δ, taken into account in Nu and in Gr is:

  • – the side, for a square;
  • δ = 0.9 D, for a disk (D = diameter);
  • δ = average length and width for a rectangle.

5.6.2. Ceiling cooling systems

image

Figure 5.5. Cooling from the top. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

For a cold surface placed in the upper part of a chamber containing the fluid, the following correlations can be used as a function of the flow regime:

  • – for math;
  • – for math.

In these relations:

  • Ra is the Rayleigh number, defined by: Ra = Gr Pr;
  • math;
  • – the characteristic dimension, δ, taken into account in Nu and in Gr is:
    • - δ = the side, for a square;
    • - δ = average length and width for a rectangle;
    • - δ = 0.9 D, for a disk (D = diameter).

5.7. Vertical cylinders subject to natural convection

image

Figure 5.6. Vertical cylinder. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

The correlations obtained for vertical cylinders are (Gebhart, 1961):

  • – for math;
  • – for math.

These relations are valid if:

  • – the ratio of the cylinder diameter, D, to its height, H, satisfies the relation (Gebhart, 1961): math;
  • – the characteristic dimension taken into account in the Nusselt and Grashof numbers is the height, H:
  • – the physical properties of the liquid are taken at the average temperature of the film, defined by: math, where θpm is the average temperature of the wall and θm is the average temperature of the fluid.

5.8. Horizontal cylinders subject to natural convection

image

Figure 5.7. Horizontal cylinder. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

In the case of horizontal cylinders subject to a natural convection, the average Nusselt number depends on the product, Gr Pr:

  • Gr Pr < 10-5: Nu = 0.49;
  • – 10-5 < Gr Pr ≤ 10-3: Nu = 0.71 (Gr Pr)1/25;
  • – 10-3 < Gr Pr ≤ 1: Nu = 1.09 (Gr Pr)0.1;
  • 1 < Gr Pr ≤ 104: Nu = 1.09 (Gr Pr)0.2;
  • – 104 < Gr Pr ≤ 109: Nu = 0.53 (Gr Pr)1/4;
  • Gr Pr > 109: Nu = 0.13 (Gr Pr)0.33.

The characteristic dimension taken into account in Nu and Gr is the external diameter of the tube, δ.

5.9. Spheres subject to natural convection

image

Figure 5.8. Immersed sphere. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

For a sphere of diameter D, we recommend using the following correlation to calculate the average Nusselt number:

math

This correlation is valid with:

  • math;
  • math;
  • Ra ≤ 1011 (Pr ≥ 0.7).

5.10. Vertical conical surfaces subject to natural convection

image

Figure 5.9. Immersed cone. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

For conical surfaces subject to a natural convection, the following correlation can be used:

math

In this correlation:

  • – the characteristic dimension to be taken into consideration when calculating the Grashof and Nusselt numbers is the height, H, of the cone;
  • math;
  • – α ≤ 12°.

5.11. Any surface subject to natural convection

For any surface, and in the absence of the possibility of estimating the natural convection heat transfer coefficient, we can use the relations valid for vertical cylinders by choosing a characteristic dimension, δ, defined by:

math

where:

Lh is the length obtained by projecting the surface considered onto a horizontal axis;

Lv is the vertical projection of the surface.

5.12. Chambers limited by parallel surfaces

image

Figure 5.10. Fluid between two horizontal plates. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

Several practical situations implement a fluid (generally air) between two plates of temperatures θp1 and θp2, respectively (θp2 > θp1). This is the case for plane solar collectors, where the absorber constitutes the hot plate (at θp2) and the glazing constitutes the second plate.

This type of configuration is also encountered in double glazing, used to minimize heat loss through windows. In the latter cases, the parallel plates are generally arranged vertically (see Figure 5.11). The hot plate can be either the plate that is in contact with the inside (as in a heated building), or that facing the outside (as in air-conditioning), according to the scenario.

image

Figure 5.11. Fluid between two vertical plates. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

For such situations, the average Nusselt number is obtained from the following general correlation:

math

where:

δ is the distance between the plates;

math

The C and n constants depend on the position of the plates and the Grashof number. They are given in Table 5.1 below.

The physical properties of the fluid separating the two plates are taken at average temperature, math:

Table 5.1. Parameters C and n (*hot plate located at the bottom)

Vertical plates
Condition on Gr C n
2 105 < Gr ≤ 1.1 107 0.071 1/3
2 103 < Gr ≤ 2 105 0.20 1/4
Gr ≤ 2 103 Nu = 1
Horizontal plates*
Condition on Gr C n
3.2 105 < Gr ≤ 107 0.075 1/3
103 < Gr ≤ 3.2 105 0.21 1/4
Gr ≤ 103 Nu = 1

5.12.1. Correlation of Hollands et al. for horizontal chambers

This is valid for gases at Ra δ< 108 and for water and liquids at Ra δ< 105. It gives the average Nu value as follows:

math

with:

math

where:

math

The exponent v indicates that the amount between the brackets is set to zero should it be negative.

5.12.2. Correlation of El-Sherbiny et al. for vertical chambers

This correlation gives the average Nu value as being the maximum of three values, Nu1, Nu2 and Nu3:

math

where:

math

This correlation is valid for:

  • – 10 ≤ Rad ≤ 2 107;
  • math.

5.13. Inclined-plane chambers

Inclined chambers are encountered in solar applications, where in order to optimize the amount of energy collected, the absorber is required to be in an inclined position (see Figure 5.12).

image

Figure 5.12. Inclined chamber. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

For such chambers, the correlation to be used depends on the inclination, α, and on the ratio, math , known as the aspect ratio.

5.13.1. For large aspect ratios and low-to-moderate inclinations

For math and 0 ≤ α ≤ 70° , we will use the following correlation (Hollands et al., 1976):

math

where:

math
math

The expressions of the type math need to be set to zero if Raδcosα < 1708

5.13.2. For lower aspect ratios and inclinations below the critical inclination

The critical inclination depends on the aspect ratio math. It is given in Table 5.2:

Table 5.2. Critical inclinations

math Critical inclination αcrit
1 25°
3 53°
6 60°
12 67°
> 12 70°

For math and math, the following correlation is recommended (Catton, 1978):

math

5.13.3. For lower aspect ratios and inclinations greater than the critical inclination

Two situations are to be considered depending on the value of αcrit determined by the aspect ratio math (see Table 5.2):

Situation 1: math

The following correlation is recommended (Catton, 1978):

math

Situation 2: math

The following correlation is recommended (Arnold et al., 1975): Nu =1+(Nuα=90° −1)sinα.

5.14. Chambers limited by two concentric cylinders

This type of chamber is generally used as a solar collector in parabolic-trough collectors of the type used in the Noor Ouarzazate power station in Morocco. They consist of two coaxial tubes, one made of metal (the inner tube), which conveys the heat-transfer fluid, and the other made of glass, which serves as a protective jacket. The fluid to be found in the annular space is generally air.

The linear flux (flux per unit length) between the absorber tube and the glass tube is given by Raithby as a function of an effective thermal conductivity λeff and a geometric influence factor, F (Raithby and Hollands, 1975). Yet the expressions proposed by Raithby introduce complexities that can be avoided.

image

Figure 5.13. Inclined chamber

Indeed, the Raithby results can be placed in a simpler form, which directly gives the average Nusselt number, and therefore the average heat transfer coefficient, h, as follows:

math

where:

math

δ is the characteristic distance, defined by: math

math

100 ≤ F Ra ≤ 1012 , where: math

The physical properties of the fluid are taken at average temperature, math:

The flux, for a tube length, L, is then obtained by the convection equation:

math

or:

math

5.15. Chambers limited by two concentric spheres

For concentric spheres of diameters D1 and D2 and of temperatures θp1 and θp2, respectively, Raithby developed an expression of the flux as a function of an effective thermal conductivity, λeff, and of the geometric influence factor, F (Raithby and Hollands, 1975).

image

Figure 5.14. Inclined chamber

Yet, as with concentric cylinders, the expressions proposed introduce complexities that can be avoided.

Indeed, the results obtained by Raithby for concentric spheres can be placed in a simpler form, which directly gives the average Nusselt number, and therefore the average heat transfer coefficient, h, as follows:

math

where:

δ is the characteristic distance, defined by: math

math

100 ≤ F Ra δ ≤ 104 , where: math

The physical properties of the fluid are determined at arithmetic average temperature:

math

The flux is then obtained by the convection equation: ϕ = hπD1D2p2 −θp1) or: math.

I.e.:

math

5.16. Simplified correlations for natural convection in air

The correlations presented above for calculating natural convection heat transfer coefficients are for general use. They are also dimensionless, enabling them to be used without any concern for problems that may be posed by units.

Specific situations can be encountered in practice, particularly when air is used as the heat-transfer fluid, as is the case with the heating or cooling of electronic components by means of natural convection. For such situations, specific correlations may be considered easier to use.

The following sections present this type of correlations. The user should be aware that these relations need to be employed with great caution with respect to the units considered for each of the magnitudes applied.

5.16.1. Vertical cylinder or plane under natural convection in air

math

where:

  • H is the height of the cylinder or plane considered, expressed in meters;
  • Tp and T are, respectively, the wall temperature and the ambient temperature far from the cylinder or the plane surface, expressed in degrees.

5.16.2. Horizontal cylinder or plane under natural convection in air

math

where:

  • H is the height of the cylinder or plane considered, expressed in meters;
  • Tp and T are, respectively, the wall temperature and the ambient temperature far from the object in question, expressed in degrees.

5.16.3. Horizontal plane under natural convection in air

The heat transfer coefficient depends on the way air flows over the hot surface:

  • – when the air is above the hot surface: math;
  • – when the air is below the hot surface: math.

Where the various parameters are defined below:

  • δ is the characteristic length of the plane considered math, expressed in meters;
    • - where A and P are the plane surface area and perimeter, respectively.
    • - Tp and T are, respectively, the wall temperature and the ambient temperature far from the object in question, expressed in degrees.

5.16.4. Sphere under natural convection in air

math

where:

  • D is the diameter of the sphere, expressed in meters;
  • Tp and T are, respectively, the wall temperature and the ambient temperature far from the sphere, expressed in degrees.

5.16.5. Circuit boards under natural convection in air

When circuit boards are cooled by natural convection in air, it is recommended to use the following correlation to determine the convective heat transfer coefficient between the boards and surrounding air:

math

where:

H is the height of the electronic board considered, expressed in meters;

Tp and T are the board temperature and the ambient temperature far from the board, respectively, expressed in degrees.

5.16.6. Electronic components or cables under natural convection in air

For small electronic components or cables under natural convection in air, it is recommended to use the following correlation to determine the convective heat transfer coefficient between the components (or cables) and surrounding air:

math

where:

  • H is the height of the electronic component or that of the cable (in meters);
  • Tp and T are, respectively, the component or cable wall temperature and the ambient temperature, expressed in degrees.

These relations are certainly helpful for rapid calculations to generate orders of magnitude of the convective heat transfer coefficients. Of course, for more accurate calculations, the dimensionless relations developed in the sections above are preferred. In addition, for calculations concerning electronic systems, the following section presents a methodology and calculations specific to heat sinks used for circuit-board systems.

5.17. Finned surfaces: heat sinks in electronic systems

During their operation, electronic components consume electric currents which, in addition to producing the desired effect, generate heat through the Joule effect. When the currents involved are significant (several amps or more), the amount of heat generated becomes so great that it can interfere with the normal operation of this component, and can even result in its disintegration, if the heat generated is not evacuated. This becomes critical when geometric constraints are imposed by design and/or ergonomic requirements, notably for embedded systems (US Department of Defense, 1992).

Thus, heat evacuation is an important parameter in the design of certain electronic systems to be used in, inter alia, telecommunications, audio equipment, avionics and computing. In a computer for example, the operation of several elements generates significant amounts of heat, particularly microprocessors (CPUs), graphics processors (GPUs), RAM, hard disks and power packs, etc. This is because these contain elementary electronic components whose operation is highly exothermic, namely, the integrated circuits that form part of most electronic assemblies. Indeed, microprocessors, such as Pentium™, Atom™ and Intelcore™, contain millions of transistors.

5.17.1. Dissipation systems

Natural convection is often sufficient in order to extract the heat produced by electronic equipment, but on the condition that it is boosted using accessories known as heat sinks (US Department of Defense, 1978). Figure 5.15 shows an example of a commercial heat sink designed to receive a transistor.

image

Figure 5.15. Transistor on heat sink assembly

When assembled on the electronic components concerned, heat sinks enable an increase in the transfer area between the electronic component considered and the neighboring air.

Heat sinks are generally composed of a metal that is a good heat conductor (copper, aluminum, etc.). In the trade, it is mainly aluminum heat sinks that are encountered.

Assembly involves screwing the electronic component onto the heat sink, making sure to insert a thermal-conductive gasket between the component and the heat sink. The latter is generally composed of a silicone or silver paste, a mica pad (the latter breaking easily, however), or a soft silicone pad (this being more practical and cleaner than paste). The role of each of these pads is to ensure thermal conduction while providing good electrical insulation (Kraus, Bar-Cohen and Wative, 2006). They can therefore be considered as insulators from an electrical perspective, but as conductors from a thermal point of view.

Yet natural convection can prove insufficient in some electronic systems. This is the case when we use performance enhancing techniques such as overclocking, which consists of exceeding the operating frequency prescribed by the manufacturer for a given component. Indeed, in such situations, operation of the electronic circuits generates such large amounts of heat that natural convection is no longer sufficient to provide the cooling needed in order to ensure stable circuit operation (US Navy, 1955).

We then use cooling systems based on forced convection, which can even draw on heat evacuation through latent heat, as is the case with spray cooling (Shicheng, Yu, Liping and Wengsuan, 2014). In such systems, in addition to the heat sinks assembled on the different components of the electronic boards, we use a fan to ensure satisfactory air circulation and, as a result, better heat transfer coefficients. This is the case for microcomputers, where a fan is often assembled inside the central processing unit box.

Although several heat sink systems exist that are based on forced convection (air, water or even liquid-nitrogen system, as in extreme cooling), above all their design falls within that of finned heat sink sizing, covered in Volume 6 of this series. In this section we will focus on systems using natural convection as the latter nevertheless remain the most frequently used for common electronic systems.

5.17.2. Thermal resistance of a heat sink

There is a very wide range of heat sinks available commercially, which differ in terms of their shape, and therefore their heat transfer areas (see Figure 5.16). Yet, in supplier catalogs, the values of the transfer areas are not presented, nor are the transfer coefficient values (see example in Table 5.3).

image

Figure 5.16. Different models of heat sinks for electronic assemblies

The heat sinks are presented with a thermal resistance. Of course, the latter depends on the geometry and the transfer area offered by a given model. Note that depending on the model chosen, the value of the resistance can vary from around 1°C/W to more than a dozen °C/W.

Table 5.3. Examples of thermal resistances in electronic heat sinks

ModelRTh (°C/W)
117.73
311.21
53.56
102.79
281.05

Let us nevertheless recall that this thermal resistance is above all determined based on the convection between the heat sink and the surrounding air, as the metal conduction resistance can be overlooked. It is of the form:

math

where:

  • h is the convection heat transfer coefficient between the heat sink and the surrounding air;
  • S is the transfer area offered by the heat sink to the surroundings.

We generally represent transfers between the heat sink and the surrounding air with an equivalent electronic circuit (see Figures 5.17 and 5.18), where:

  • TJ is the junction temperature between the electronic component considered and the case representing the heat sink; it is the temperature of the component;
  • Ta is the ambient air temperature.
image

Figure 5.17. Simplified electrical representation of heat transfer between a heat sink and surrounding air

With all thoroughness, each of the thermal resistances interposed between the electronic component and the air should be taken into account.

image

Figure 5.18. Detailed electrical representation of heat transfer between an electronic component and the surroundings. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

We then take into account the following thermal resistances:

  • RJC: between the electronic component (the junction) and the case. It is a conductive resistance that depends on the thermal-conductive gasket used. Its value can be deduced from the thermal-conductor thickness, eTC, its contact surface area, SC, and its thermal conductivity, λTC:
    math
  • RCS: between the case and the heat sink. Here too we are looking at a conductive resistance, the value of which depends not only on the thickness and thermal conductivity of the case, but also the thickness and thermal conductivity of the electrical insulator used;
  • RSA: between the heat sink and the air: math.

As the resistances RJC and RCS are generally the lowest (conductive resistances), optimization of the overall resistance will require minimizing RSA; the product hS needs therefore to be maximized.

5.18. Optimizing the thermal resistance of a heat sink

The choice of heat sink geometry determines the transfer area between the latter and the air. We generally opt for finned surfaces (see Figure 5.19).

For a vertical-fin heat sink, the natural convection heat transfer coefficient is determined based on Bar-Cohen and Rohsenow correlations for the two situations usually encountered: constant wall temperature and constant flux at the wall (Bar-Cohen and Rohsenow, 1984).

image

Figure 5.19. Details of the heat-sink fins

In general, the L and math dimensions of the heat sink are determined by the space restrictions on the electronic board considered. When L is known, we can determine the maximum number of fins that can be put in place, namely:

math

However, the number of fins to be inserted effectively conditions the heat transfer coefficient, h. Indeed on a given base surface (A = math x L) placing a large number of fins (n large) will lead to a low spacing δ. This means that we will need a large stacking of fins, leading to a poor air circulation and therefore to a lower heat transfer coefficient.

If, conversely, we opt for a smaller number of fins, air circulation will be facilitated and, as a result, the heat transfer coefficient will be larger.

Yet the number of fins, n, conditions the transfer area: S = 2n math H.

Thus, when S increases, h is lower and conversely, when S is lower h is greater. The optimum heat transfer area is thus defined by seeking a compromise between the number of fins to be inserted on a given base surface (math x L), and the corresponding heat transfer coefficient, h.

It follows that δ admits an optimum that makes the hS product maximal.

Yet the different experiments conducted in the field have shown that this optimum depends on the natural convection heat transfer coefficient between the heat sink and the air.

5.18.1. Determining the heat-sink/air heat transfer coefficient

In the case of heat transfer between a finned heat sink and air, we consider that energy transfers essentially via the fins and that the temperature of the latter is constant in steady state. The Nusselt number is obtained from the correlation presented below (Kraus, Bar-Cohen and Wative, 2006):

math

where the various parameters are defined as follows:

  • math;
  • h is the natural convection heat transfer coefficient between the heat sink and the surrounding air;
  • δ is the spacing between two fins;
  • E𝓁 is the Elenbaas number, defined by: math;
  • math;
  • λ, ρ, μ, β: respectively, the air thermal conductivity, its density, its viscosity and its thermal expansion coefficient math;
  • g is the acceleration of gravity;
  • – the air physical properties are taken at: math;
  • T and Ta are the fin temperature and the air temperature, respectively.

5.18.2. Calculating the optimum spacing between fins

The optimum space between the fins is determined by Bar-Cohen and Rohsenow as follows:

math

where Ra𝓁 is the Rayleigh number, calculated taking math as the characteristic length, namely:

math

5.18.3. Practical expression

It is also demonstrated that the product hδ* depends only on air heat conductivity, namely:

math

5.18.4. Calculating the evacuated heat flux

Once the heat transfer coefficient and the optimum number of fins are known, the evacuated flux is given by:

math

Or, substituting for δ* and n* and simplifying overlooking e in favor of δ.

math

5.18.5. Implementation algorithm

For calculations (Ellison, 1984), we generally proceed in accordance with the flowchart presented in Figure 5.20.

image

Figure 5.20. Algorithm to optimize the surface of a heat sink

5.18.6. Illustration: optimum design of a heat sink

In the design of the motherboard of a central processing unit which should receive a microprocessor, a space with dimensions math has been reserved for assembly of the microprocessor/heat-sink system. By choice, it is decided to ensure cooling of this microprocessor by natural convection. To this end, we select a commercially available vertical-fin heat sink with thickness e and dimensions math. The microprocessor’s optimum operating power and temperature are respectively ϕ0 and T*.

Questions
  1. 1) Calculate the optimum spacing to be left between the fins in order to ensure satisfactory evacuation of the heat generated by the microprocessor’s operation.
  2. 2) Deduce therefrom the optimum number of fins needing to be placed on the heat sink.
  3. 3) Calculate the convection heat transfer coefficient between the heat sink and the surrounding air.
  4. 4) Will the microprocessor operate correctly?

    Data:

    • L = 5 cm
    • math = 3 cm
    • H = 6 cm
    • e = 2 mm
    • T* = 93 °C
    • 𝛟0 = 10 W
    • g = 9.81 m/s2
    • Air: Ta = 27 °C

Table 5.4. Physical properties of air at different temperatures

(°C) (kg. m-3) Cp (J.kg-1.°C-1) (W.m-1.°C-1) Pa sec
0 1.292 1,006 0.0242 1.72E-05
20 1.204 1,006 0.0257 1.81E-05
40 1.127 1,007 0.0272 1.90E-05
60 1.059 1,008 0.0287 1.99E-05
80 0.999 1,010 0.0302 2.09E-05
100 0.946 1,012 0.0318 2.18E-05
120 0.898 1,014 0.0333 2.27E-05
Solutions

1) Calculating the optimum spacing between fins

The optimum fin spacing is given by: math.

Where math is defined using math.

2) Optimum number of fins to be placed on the heat sink

The optimum number of fins to be planned for on the heat sink is given by:

math

3) Convection heat transfer coefficient between the heat sink and the air

math

4) Microprocessor operation

During the course of its operation at T*, the microprocessor will generate a power, ϕ0, which will need to be evacuated in order to ensure smooth operation. We then calculate the power that the available heat sink would be able to evacuate, ϕ*, and we compare it to ϕ0.

The heat sink will make it possible to evacuate a flux given by:

math

where: math.

5.19. Optimum circuit-board assembly

In this section we will be considering the assembly of a set of circuit boards on a rack placed in a box (see Figure 5.21).

We will assume that all of the boards are the same size, and that the rack gathering them together is of length L and width math. The spacing between the different boards is constant (δ), but it needs to be such that heat evacuation is optimum. We will also assume that each board releases a constant flux density, φ.

image

Figure 5.21. Positioning of circuit boards on a rack. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

Under these conditions, the temperature of the boards is not uniform; rather it increases with height, reaching its maximum value at H. The heat transfer occurs by natural convection.

The question here is, how many boards can we place on the board whilst assuring optimum heat evacuation? In other words, what is the optimum spacing, δ?

5.19.1. Calculating the optimum spacing between electronic boards

This spacing was determined through experimentation by analyzing different situations with different electronic boards and different spacings. This work led to the correlation of Bar-Cohen and Rohsenow:

math

where Raφ is the Rayleigh number relating to the flux density, φ, and using math as the characteristic length:

math

Hence the optimum number of boards of thickness e to be assembled on a rack of length L is:

math

5.19.2. Heat transfer coefficient between electronic boards and air

The Nusselt number that determines heat exchanges between electronic boards and neighboring air has also been established in experiments conducted by Bar-Cohen and Rohsenow:

math

where:

  • math is the modified Elenbaas number, defined based on the flux density, φ
    math
  • Raφ is the modified Rayleigh number, defined based on the flux density, φ
    math
  • h is the natural convection heat transfer coefficient between board and air;
  • δ is the spacing between electronic boards;
  • – λ, ρ, μ are the air thermal conductivity, density and viscosity, respectively;
  • β is the air thermal expansion coefficient: math;
  • – g is the acceleration of gravity.

Air physical properties are taken at math and Ta are the temperatures of the board and the air, respectively.

5.19.3. Calculating the evacuated heat flux

Knowing the heat transfer coefficient and the optimum number of electronic boards to place on the rack, we calculate the evacuated heat flux:

math

or alternatively, by overlooking e in favor of math.

5.19.4. Implementation algorithm

Often, we know the thermal flux released by the electronic boards, but we do not know the temperature that will be reached by the boards during operation. As a result, we do not directly know the physical properties of air, which are fundamental in determining the Rayleigh number. In a situation such as this, calculations are performed iteratively, as shown in the flowchart in Figure 5.22.

image

Figure 5.22. Optimization of electronic board installation on racks

5.19.5. Illustration: optimum evacuation of heat generated by electronic boards

Putting in place an electronic surveillance system requires the assembly of k identical circuit boards on a rack that must be placed in a box. The rack presents length L and width math.

For ease of installation’s sake, the spacing between the different boards, δ, will be uniform. Moreover, to avoid eventual fan noise, heat evacuation will only be by natural convection. The system should thus be designed so that this heat evacuation is sufficient to ensure safe and correct operation. Under these conditions the power generated by each board is ϕ*.

Questions
  1. 1) Calculate the optimum spacing to be left between the boards.
  2. 2) What number of boards do you propose as a limit for k?

    Data:

    • ϕ* = 67 W
    • k = 7
    • L = 40 cm
    • math = 35 cm
    • H = 20 cm
    • e = 8 cm
    • g = 9.81 m/s2
    • AirTa = 30 °C

Table 5.5. Physical properties of air between 30 and 60°C

(°C) (kg. m-3) Cp (J.kg-1.°C-1) (W.m-1.°C-1) Pa sec
30 1.173 1,006.7 0.0265 1.86E-05
35 1.135 1,006.9 0.0269 1.88E-05
40 1.127 1,007.0 0.0272 1.90E-05
45 1.065 1,007.3 0.0277 1.93E-05
50 1.086 1,007.5 0.0281 1.95E-05
55 1.066 1,007.8 0.0284 1.97E-05
60 1.059 1,008.0 0.0287 1.99E-05
Solutions

1) Calculating the optimum spacing to be left between the boards

As the operating temperature is not known, we will proceed by successive iterations, starting from T0 = Ta = 30°C.

The results of these iterations are presented in Table 5.16. It follows that, for the present case, T0 does not have much influence on the calculations, where math.

Table 5.6. Calculating the operating temperature

T°C Tave °C ρ kg. m-3 CpJ.kg-1.°C-1 λ W.m-1.°C-1 μ Pa sec ß°K-1 Pr Rai δ*m n* h* W/m-°C T°C
30 30 1.173 1,006.7 0.0265 1.86E-05 3.30E-03 0.7066 2.47E+10 1.87E-03 4.89 375.88 31.27
40 35 1.135 1,006.9 0.0269 1.88E-05 3.19E-03 0.7037 2.15E+10 1.94E-03 4.88 360.95 31.33
50 40 1.127 1,007 0.0272 1.90E-05 3.10E-03 0.7034 1.98E+10 1.98E-03 4.88 353.72 31.35
60 45 1.065 1,007.3 0.0277 1.93E-05 3.00E-03 0.7018 1.63E+10 2.08E-03 4.87 333.13 31.44
70 50 1.086 1,007.5 0.0281 1.95E-05 2.92E-03 0.6992 1.59E+10 2.09E-03 4.87 334.01 31.43

We will retain T* = 31.3°C.

2) Number of boards

Based on the calculations above, we will retain k = 4 with a minimum spacing of 2 mm.

5.20. Superimposed forced and natural convections

In general, when a fluid is subject to a forced convection, heat exchanges imposed by the fluid motion are so significant that the effects of natural convection are negligible. This is the case, in particular, for turbulent flow. When the flow is laminar, however, the contribution of natural convection can play an important role in overall heat transfer, particularly when the velocity and temperature profiles of forced convection are influenced by the effects of natural convection (Metais and Eckert, 1964). In this case, the convective heat transfer coefficients are substantially influenced and the expressions presented in the above sections are no longer applicable.

Thus, the correlations to be used in this case need to take into account both contributions: that of natural convection combined with the contribution of forced convection.

Moreover, given that natural convection depends on gravitational force, the correlations also need to consider the orientation of the pipe used (vertical or horizontal).

5.20.1. Vertical-tube scenario: Martinelli-Boelter correlation

Martinelli and Boelter studied the scenario of natural convection superimposed onto a forced convection in vertical tubes (see Figure 5.23).

image

Figure 5.23. Natural and forced convections in a vertical tube. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

For situations where the fluid velocity is low, they developed the following correlation (Martinelli and Boelter, 1942):

math

where:

Table 5.7. Factors math and math for the Martinelli-Boelter correlation

Z 0 0.1 0.2 0.3 0.4 0.5 1.0 1.5 1.7 1.8 1.9 2
F1 1 0.997 0.993 0.990 0.985 0.978 0.912 0.770 0.675 0.610 0.573 0
F2 1 0.952 0.910 0.869 0.828 0.787 0.588 0.403 0.320 0.272 0.212 0

5.20.2. Horizontal-tube scenario: Proctor-Eubank correlation

The case where natural convection is superimposed onto forced convection in horizontal tubes was explored by Eubank and Proctor (see Figure 5.24).

It was demonstrated that this situation is only significant in the case of large temperature variations, Δθ, between the wall and the fluid; only in this case do the differences in density become sufficiently significant to modify the laminar currents of forced convection.

Thus, only under such conditions are the heat transfer coefficients affected by a perceptible influence of natural convection on forced convection.

image

Figure 5.24. Natural and forced convections in a horizontal tube. For a color version of this figure, see www.iste.co.uk/benallou/energy3.zip

The following correlation is then proposed (Eubank and Proctor, 1951):

math

where:

Gre is the Grashof number corresponding to the temperature difference, Δθe, between the wall and the fluid at the tube inlet:

math

All of the physical properties are taken at the fluid’s average temperature: θm;

5.20.3. Cylinders, disks or spheres in rotation

In the case of bodies rotating at low velocity, natural convection is superimposed onto laminar flow transfer. Indeed, the rotation of the body induces a forced convection which is laminar in nature for low rotational velocities. Moreover, when the difference between the temperature of the medium in which the body is rotating and that of the wall of the latter is fairly significant, the differences in density of the fluid constituting the medium become noticeable and consequently have an influence on heat transfer.

5.20.3.1. Cylinder in rotation

image

Figure 5.25. Rotating cylinder

For a cylinder with a wall temperature of θp, rotating at a velocity, ω, in a fluid (air, gas or any liquid) at average temperature, θm, the average Nusselt number is given by:

math

where:

  • math, where ω is the rotational velocity in revolutions per second;
  • NuD is the Nusselt number related to the diameter, math ;
  • math.

5.20.3.2. Disk in rotation

image

Figure 5.26. Disk in rotation

For a disk of diameter D, at wall temperature θp, rotating at a velocity, ω, in the air, which is at average temperature θm, the average heat transfer coefficient can be estimated by:

math

where:

  • math, where ra is the rotational velocity (revolutions per second);
  • math;
  • math.

5.20.3.3. Sphere in rotation

Although rarely encountered in practice, the case of spheres rotating in media at a different temperature is considered in this section.

Consider a sphere of temperature θp, rotating at a velocity, ω, in a fluid at average temperature θm (see Figure 5.27).

image

Figure 5.27. Sphere in rotation

The average heat transfer coefficient depends on the Reynolds range considered:

  • – for Re ≤ 5 102: Nu = 0.43 Re0.5 Pr0.4+ 0.1 (GrPr)0.3;
  • – for 5 104 < Re ≤ 7 105: Nu = 0.066 Re0.67 Pr0.4+ 0.08 (GrPr)0.1.

In these correlations:

  • math, where ω is the rotational velocity in revolutions per second;
  • math;
  • math.