Learning Unit 4 - Convection Heat Transfer
Learning Unit 4 - Convection Heat Transfer
Learning outcomes
After studying this unit, you should be able to
Convection is the mechanism of heat transfer through a fluid in the presence of bulk
fluid motion. Convection is a function of the temperature difference between the
surface and the fluid and the heat transfer coefficient. The heat transfer coefficient is
not a physical property of the material but depends on a few parameters including fluid
properties as well as the nature of the fluid motion.
Convection is classified as natural (or free) and forced convection, depending on how
the fluid motion is initiated. In forced convection, the fluid is forced to flow over a
surface or in a pipe by external means such as a pump or a fan. In natural convection,
any fluid motion is caused by natural means such as the buoyancy effect, which
manifests itself as the rise of warmer fluid and the fall of the cooler fluid. Convection
is also classified as external and internal, depending on whether the fluid is forced to
flow over a surface or in a pipe.
Open Rubric
dimensional analysis, which reduces the number of controlling parameters to a few
non-dimensional groupings.
Convection is
• extremely diverse
• several parameters involved (fluid properties, geometry, nature of flow, phases,
etc)
• systematic approach required
• classify flows into certain types, based on certain parameters
• identify parameters governing the flow, and group them into meaningful non-
dimensional numbers
• need to understand the physics behind each phenomenon
Conduction and convection both require the presence of a material medium, but
convection requires fluid motion. Convection involves fluid motion as well as heat
conduction. Heat transfer through a solid is always by conduction. Heat transfer
through a fluid is by convection in the presence of bulk fluid motion and by conduction
in the absence of it. Therefore, conduction in a fluid can be viewed as the limiting case
of convection, corresponding to the case of quiescent fluid (Fig. 4.1).
Fig. 4.1: Heat transfer from a hot surface to the surrounding fluid by convection
and conduction (Cengel and Ghajar, 2020).
The fluid motion enhances heat transfer, since it brings warmer and cooler chunks of
fluid into contact, initiating higher rates of conduction at a greater number of sites in a
fluid. The rate of heat transfer through a fluid is much higher by convection than it is
by conduction. In fact, the higher the fluid velocity, the higher the rate of heat transfer.
2
Convection is a complex physical phenomenon governed by several
parameters. Experience shows that convection heat transfer strongly depends on
the fluid properties dynamic viscosity μ, thermal conductivity k, density ρ, and specific
heat cp, as well as the fluid velocity V. It also depends on the geometry and the
roughness of the solid surface, in addition to the type of fluid flow (such as being
streamlined or turbulent). Thus, we expect the convection heat transfer relations to be
rather complex because of the dependence of convection on so many variables.
Despite the complexity of convection, the rate of convection heat transfer is observed
to be proportional to the temperature difference and is conveniently expressed by
Newton’s law of cooling as
Or
where,
From its units, the convection heat transfer coefficient h can be defined as the rate of
heat transfer between a solid surface and a fluid per unit surface area per unit
temperature difference.
Evaluation of the convective or convection heat transfer coefficient then completes the
parameters necessary for heat transfer calculations. The convective heat transfer
coefficient is evaluated in some limited cases by mathematical analytical methods and
in most cases by experiments.
• flow field
• temperature field in fluid
• heat transfer coefficient, h.
How do we determine h?
Consider the process of convective cooling, as we pass a cool fluid past a heated wall.
This process is described by Newton’s law of cooling. Near any wall a fluid is subject
to the no slip condition; that is, there is a stagnant sub-layer. A consequence of the
3
no-slip condition is that all velocity profiles must have zero values with respect to the
surface at the points of contact between a fluid and a solid surface (Fig. 4.2).
Fig. 4.2: A fluid flowing over a stationary surface comes to a complete stop at the
surface because of the no-slip condition (Cengel and Ghajar, 2020).
Since there is no fluid motion in this layer, heat transfer is by pure conduction in this
region, and it can be expressed as
𝜕𝜕𝜕𝜕
𝑞𝑞̇ 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 = 𝑞𝑞̇ 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 = −𝑘𝑘𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓 𝑑𝑑𝑑𝑑 ⃒𝑦𝑦=0 (𝑊𝑊 ⁄𝑚𝑚2 ) (4.3)
where T represents the temperature distribution in the fluid and (∂T/∂y)y=0 is the
temperature gradient at the surface and it depends on the whole fluid motion, and both
fluid flow and heat transfer equations are needed. Above the sub-layer is a region
where viscous forces retard fluid motion; in this region some convection may occur,
but conduction may well predominate. A careful analysis of this region allows us to
use our conductive analysis in analyzing heat transfer. This is the basis of our
convective theory.
At the wall, heat is then convected away from the surface because of fluid motion.
Convection heat transfer from a solid surface to a fluid is merely the conduction heat
transfer from the solid surface to the fluid layer adjacent to the surface. We can
equate Eqs. 4.1 and 4.3 for the heat flux to obtain for the determination of the
convection heat transfer coefficient when the temperature distribution within the fluid
is known.
−𝑘𝑘𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓 (𝜕𝜕𝜕𝜕⁄𝜕𝜕𝜕𝜕)𝑦𝑦=0
ℎ= (𝑊𝑊 ⁄𝑚𝑚2 . 𝐾𝐾) (4.4)
𝑇𝑇𝑠𝑠 −𝑇𝑇∞
This expression shows that to determine h, we must first determine the temperature
distribution in the thin fluid layer that coats the wall.
In convection heat transfer, the key unknown is the heat transfer coefficient. Also, it is
common practice to nondimensionalize the governing equations and combine the
variables, which group together into dimensionless
numbers in order to reduce the number of total variables.
4
From Eq. 4.4, we obtain the following equation in terms of dimensionless parameters:
𝑘𝑘 (𝑇𝑇 −𝑇𝑇 ) 𝜕𝜕𝜕𝜕 𝑘𝑘 𝜕𝜕𝜕𝜕
ℎ = − 𝐿𝐿𝑓𝑓 �(𝑇𝑇∞−𝑇𝑇 𝑠𝑠 )� 𝑑𝑑𝑑𝑑 ⃒𝑦𝑦=0 = + 𝐿𝐿𝑓𝑓 𝑑𝑑𝑑𝑑 ⃒𝑦𝑦=0 (4.5)
𝑐𝑐 𝑠𝑠 ∞ 𝑐𝑐
Inspection of this equation suggests that the appropriate dimensionless form of the
heat transfer coefficient is the so-called Nusselt number, Nu, defined by
𝒉𝒉𝑳𝑳𝒄𝒄
𝑵𝑵𝑵𝑵 = (4.6)
𝒌𝒌
where k is the thermal conductivity of the fluid and Lc is the characteristic length.
Fig. 4.3: Heat transfer through a fluid layer of thickness L and temperature difference
ΔT.
Heat transfer through the fluid layer is by convection when the fluid involves some
motion and by conduction when the fluid layer is motionless. Heat flux (the rate of heat
transfer per unit surface area) in either case is
which is the Nusselt number. This means that the Nusselt number represents the
enhancement of heat transfer through a fluid layer because of convection relative to
conduction across the same fluid layer. The larger the Nusselt number, the more
effective the convection. A Nusselt number of Nu = 1 for a fluid layer represents heat
transfer across the layer by pure conduction.
A. Based on geometry:
External flow: The flow of an unbounded fluid over a surface such as a plate, a wire,
or a pipe.
Internal flow: The flow in a pipe or duct if the fluid is completely bounded by solid
surfaces.
Forced flow: A fluid is forced to flow over a surface or in a pipe by external means
such as a pump or a fan.
Natural flow: Fluid motion is due to natural means such as the buoyancy effect, which
manifests itself as the rise of warmer (and thus lighter) fluid and the fall of cooler (and
thus denser) fluid.
Laminar/turbulent
Laminar flow: The highly ordered fluid motion characterized by smooth layers of fluid.
The flow of high-viscosity fluids such as oils at low velocities is typically laminar.
Turbulent flow: The highly disordered fluid motion that typically occurs at high
velocities and is characterized by velocity fluctuations. The flow of low-viscosity fluids
such as air at high velocities is typically turbulent.
Transitional flow: A flow that alternates between being laminar and turbulent.
Compressible/Incompressible flow
Compressible flow: If the density of fluid changes during flow (e.g., high-speed gas
flow)
6
F. Based on velocity distribution
One/two/Multi-dimensional flow
A flow is said to be one-, two-, or three-dimensional if the flow velocity varies in one,
two, or three dimensions, respectively.
G. Based on viscosity and regions of flow
Inviscid flow regions: In many flows of practical interest, there are regions (typically
regions not close to solid surfaces) where viscous forces are negligibly small
compared to inertial or pressure forces.
Fig. 4.4 shows the velocity distribution at various distances from the leading edge of a
flat plate. The x-coordinate is measured along the plate surface from the leading edge
of the plate in the direction of the flow, and y is measured from the surface in the
normal direction. The fluid approaches the plate in the x-direction with a uniform
velocity V, which is practically identical to the free-stream velocity over the plate away
from the surface.
Fig. 4.4: The development of the boundary layer for flow over a flat plate, and the
different flow regimes (Cengel and Ghajar, 2020).
7
Due to no-slip condition, the velocity of the particles in the first fluid layer adjacent to
the plate becomes zero. This motionless layer slows down the particles of the
neighbouring fluid layer because of friction between the particles of these two adjoining
fluid layers at different velocities and so on. Thus, the presence of the plate is felt up
to some normal distance δ from the plate beyond which the free-stream velocity
remains essentially unchanged. As a result, the x-component of the fluid velocity,
u, varies from 0 at y = 0 to nearly V at y = δ (Fig. 4.5).
Fig. 4.5: The development of a boundary layer on a surface is due to the no-
slip condition and friction (Cengel and Ghajar, 2020).
The region of the flow above the plate bounded by δ in which the effects of the viscous
shearing forces caused by fluid viscosity are felt is called velocity boundary layer. The
flow region near the plat where the velocity of the fluid is decreased by viscous forces
is called boundary layer. The boundary layer thickness, δ, is typically defined as the
distance y from the surface at which u = 0.99V. The distance from the plate at which
the velocity reaches 99% of the free-stream velocity is arbitrarily designated as
boundary layer thickness, and the region beyond this point is called free-stream
region.
In Fig. 4.3, initially, the flow in the boundary layer is completely laminar. The boundary
layer thickness grows with increasing distance from the leading edge, and at some
critical distance xcr, the inertial effects become sufficiently large compared to the
viscous damping action that small disturbances in the flow begin to grow. As these
disturbances become amplified, the regularity of the viscous flow is disturbed and a
transition from laminar to turbulent flow takes place. In the turbulent-flow region,
macroscopic chunks of fluid move across streamlines and vigorously transport thermal
energy as well as momentum.
From the leading edge onward, a region develops in the flow in which viscous forces
cause the fluid to slow down. These viscous forces depend on the shear stress 𝜏𝜏. In
flow over a flat plate, the fluid velocity parallel to the plate can be used to define the
stress as, i.e., shear stress for most fluids is proportional to the velocity gradient
(∂u/∂y), and the shear stress at the wall surface:
𝜕𝜕𝜕𝜕
𝜏𝜏𝑤𝑤 = 𝜇𝜇 𝑑𝑑𝑑𝑑 ⃒𝑦𝑦=0 (𝑁𝑁⁄𝑚𝑚2 ) (4.10)
where the constant of proportionality μ is the dynamic viscosity of the fluid (kg/m⋅s or
N⋅s/m2, or Pa⋅s).
8
A more practical approach in external flow is to relate τw to the upstream velocity V
as
𝜌𝜌𝑉𝑉 2
𝜏𝜏𝑤𝑤 = 𝐶𝐶𝑓𝑓 (𝑁𝑁⁄𝑚𝑚2 ) (4.11)
2
Once the average friction coefficient over a given surface is available, the friction
force over the entire surface is determined from
𝝆𝝆𝑽𝑽𝟐𝟐
𝑭𝑭𝒇𝒇 = 𝑪𝑪𝒇𝒇 𝑨𝑨𝒔𝒔 (𝑵𝑵) (4.12)
𝟐𝟐
Fig. 4.6 shows the development of a thermal boundary layer. A thermal boundary layer
develops in a similar manner to the velocity boundary layer if there is a difference
between the fluid free stream temperature and the temperature of the plate. A thermal
boundary layer develops when a fluid at a specified temperature flows over a surface
that is at a different temperature.
Fig. 4.6: Thermal boundary layer on a flat plate (the fluid is hotter than the plate
surface) (Cengel and Ghajar, 2020).
In thermal boundary, the flow region over the surface in which the temperature
variation in the direction normal to the surface is significant. The thickness of the
thermal boundary layer δt at any location along the surface is defined as the distance
from the surface at which the temperature difference T−Ts equals 0.99(T∞− Ts).
The thickness of the thermal boundary layer increases in the flow direction since the
effects of heat transfer are felt at greater distances from the surface further
downstream.
The shape of the temperature profile in the thermal boundary layer dictates the
convection heat transfer between a solid surface and the fluid flowing over it.
9
4.4.3 Prandtl Number
The relative thickness of the velocity and the thermal boundary layers is best described
by the dimensionless parameter Prandtl number
The Prandtl numbers of gases are about 1, which indicates that both momentum and
heat dissipate through the fluid at about the same rate. Heat diffuses very quickly in
liquid metals (Pr ≪ 1) and very slowly in oils (Pr ≫ 1) relative to
momentum. Consequently, the thermal boundary layer is much thicker for liquid
metals and much thinner for oils relative to the velocity boundary layer.
The physical meaning of Prandtl number is that it may be seen to be a ratio reflecting
the ratio of the rate that viscous forces penetrate the material to the rate that thermal
energy penetrates the material. Consequently, the Prandtl number is proportional to
the rate of growth of the two boundary layers.
In laminar boundary layer, smooth streamlines and highly ordered motion of the fluid
occurs. In contrast, fluid motion in turbulent boundary layer experience velocity
fluctuations and highly disordered motion.
The transition from laminar to turbulent flow does not occur suddenly; rather, it occurs
over some region in which the flow fluctuates between laminar and turbulent flows
before it becomes fully turbulent. Most flows encountered in practice are turbulent.
Laminar flow is encountered when highly viscous fluids such as oils flow in small pipes
or narrow passages.
Typical average velocity profiles in laminar and turbulent flow are also given in Fig.
4.4. The turbulent boundary layer can be considered to consist of four regions,
characterized by the distance from the wall. The very thin layer next to the wall where
viscous effects are dominant is the viscous sublayer. The velocity profile in this layer
is very nearly linear, and the flow is streamlined. Next to the viscous sublayer is the
buffer layer, in which turbulent effects are becoming significant, but the flow is still
dominated by viscous effects. Above the buffer layer is the overlap layer, in which the
turbulent effects are much more significant, but still not dominant. Above that is the
turbulent layer, in which turbulent effects dominate over viscous effects.
where V is the upstream velocity (equivalent to the free-stream velocity for a flat plate),
Lc is the characteristic length of the geometry, and ν = μ/ρ is the kinematic viscosity of
the fluid.
The Reynolds number at which the flow becomes turbulent is called the critical
Reynolds number. The value of the critical Reynolds number is different for different
geometries and flow conditions. For flow over a flat plate, the generally accepted value
of the critical Reynolds number is Recr = Vxcr /v = 5 ×105, where xcr is the distance from
the leading edge of the plate at which transition from laminar to turbulent flow occurs.
In this section, it is assumed that the student is familiar with flow governing equations.
For clarity, the flow equations are stated.
The continuity equation:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
+ 𝜕𝜕𝜕𝜕 = 0 (4.15)
𝜕𝜕𝜕𝜕
This is the conservation of mass relation in differential form, which is also known as
the continuity equation or mass balance for steady two-dimensional flow of a fluid with
constant density.
This is the relation for the momentum balance in the x-direction, and it is known as
the x-momentum equation.
This equation states that the net energy convected by the fluid out of the control
volume is equal to the net energy transferred into the control volume by
heat conduction.
Recall that the boundary layer thickness is defined as the distance from the surface
for which u/V = 0.99. From Table 6–3 (Cengel and Ghajar, 2020) that the value of η
corresponding to u/V = 0.99 is η = 4.91. Substituting η = 4.91 and y = δ into the
definition of the similarity variable
11
𝑉𝑉
𝜂𝜂 = 𝑦𝑦�𝜈𝜈𝜈𝜈 (4.18)
since Rex = Vx/v, where x is the distance from the leading edge of the plate.
Unlike the boundary layer thickness, wall shear stress and the skin friction coefficient
decrease along the plate as x−1/2.
Knowing the velocity profile, we can solve the energy equation for temperature
distribution for the case of constant wall temperature Ts. The local convection
coefficient and Nusselt number can be expressed as
and
𝒉𝒉𝒙𝒙 𝒙𝒙 𝟏𝟏⁄𝟐𝟐
𝑵𝑵𝑵𝑵𝒙𝒙 = = 𝟎𝟎. 𝟑𝟑𝟑𝟑𝟑𝟑𝐏𝐏𝐏𝐏𝟏𝟏⁄𝟐𝟐 𝐑𝐑𝐑𝐑𝐱𝐱 𝐏𝐏𝐏𝐏 > 𝟎𝟎. 𝟔𝟔 (4.22)
𝒌𝒌
The Nu x values obtained from this relation agree well with measured values.
It is also determined that δ/δt ≅ Pr1/3 hence the thermal boundary layer
thickness becomes
𝜹𝜹 𝟒𝟒.𝟗𝟗𝟗𝟗𝟗𝟗
𝜹𝜹𝒕𝒕 = = (4.23)
𝐏𝐏𝐏𝐏𝟏𝟏⁄𝟑𝟑 𝐏𝐏𝐏𝐏𝟏𝟏⁄𝟑𝟑 �𝑹𝑹𝑹𝑹𝒙𝒙
It should be noted that these relations are valid only for laminar flow over an
isothermal flat plate. Also, the effect of variable properties can be accounted for by
evaluating all such properties at the film temperature defined as Tf=(Ts+T∞)/2.
The friction coefficient for a given geometry can be expressed in terms of the Reynolds
number Re and the dimensionless space variable alone (instead of being expressed
in terms of x, L, V, ρ, and μ). This is a very significant finding, and it shows the value
of nondimensionalized equations.
12
Relations exist which state that for a given geometry, the friction coefficient can be
expressed as a function of Reynolds number alone, and the Nusselt number as a
function of Reynolds and Prandtl numbers alone.
Experiments have been conducted which report friction and heat transfer coefficient
measurements conveniently in terms of Reynolds and Prandtl numbers. The
experimental data for heat transfer is often represented with reasonable accuracy by
a simple power-law relation of the form
𝐍𝐍𝐍𝐍 = 𝐂𝐂𝐑𝐑𝐑𝐑𝐦𝐦
𝐋𝐋 𝐏𝐏𝐏𝐏
𝐧𝐧
(4.25)
where m and n are constant exponents (usually between 0 and 1), and the value of
the constant C depends on geometry.
Consider the parallel flow of a fluid over a flat plate of length L in the flow direction, as
shown in Fig. 4.7. The x-coordinate is measured along the plate surface from the
leading edge in the direction of the flow. The fluid approaches the plate in the x-
direction with a uniform velocity V and temperature T∞. The flow in the velocity
boundary layers starts out as laminar, but if the plate is sufficiently long, the flow
becomes turbulent at a distance xcr from the leading edge where the Reynolds number
reaches its critical value for transition.
Fig. 4.7: Laminar and turbulent regions of the boundary layer during flow over a flat
plate.
The transition from laminar to turbulent flow depends on the surface geometry, surface
roughness, upstream velocity, surface temperature, and the type of fluid, among other
things, and is best characterized by the Reynolds number.
The Reynolds number at a distance x from the leading edge of a flat plate is expressed
as
13
𝝆𝝆𝝆𝝆𝝆𝝆 𝑽𝑽𝑽𝑽
𝐑𝐑𝐑𝐑𝐱𝐱 = 𝝁𝝁
= 𝝂𝝂
(4.26)
Since the heat transfer coefficient is a direct function of the temperature gradient next
to the wall, the physical variables on which it depends can be expressed as follows:
As previously noted, since the function is dependent on several parameters, the heat
transfer coefficient is usually expressed in terms of correlations involving pertinent
non-dimensional numbers.
Viscous force provides the dampening effect for disturbances in the fluid. If dampening
is strong enough, we have laminar flow. Otherwise, instability will lead to turbulent flow
and critical Reynolds number.
For forced convection, the heat transfer correlation can be expressed as Nu = f (Re,
Pr).
The local Nusselt number at a location x for laminar flow over a flat plate for the case
of constant wall temperature was determined as:
ℎ𝑥𝑥 𝑥𝑥 1⁄3
𝑁𝑁𝑁𝑁𝑥𝑥 = = 0.332Re0.5
x Pr Pr > 0.6, 𝑅𝑅𝑅𝑅𝑥𝑥 < 5 x 105 (4.27)
𝑘𝑘
where Nux = hxx/k; Rex = ρVx/μ, Pr =cpμ/k and hx is the local heat transfer coefficient.
To put Eq. 4.27 back into dimensional form, we replace the Nusselt number by its
equivalent, hxx/k and take the x/k to the other side:
14
We can see that the convective coefficient decreases with x½. Also, as the boundary
layer thickens, the convection coefficient decreases.
If we are interested in the total heat loss from a surface, rather than the temperature
at a point, then we may well want to know something about average convective
coefficients, hL.
The desire is to find a correlation that provides an overall heat transfer rate:
𝐿𝐿
𝑄𝑄̇ = ℎ𝐿𝐿 𝐴𝐴(𝑇𝑇𝑠𝑠 − 𝑇𝑇∞ ) = ∫ ℎ𝑥𝑥 (𝑇𝑇𝑠𝑠 − 𝑇𝑇∞ )𝑑𝑑𝑑𝑑 = ∫0 ℎ𝑥𝑥 (𝑇𝑇𝑠𝑠 − 𝑇𝑇∞ )𝑑𝑑𝑑𝑑 (4.31)
where hx and hL, refer to local and average convective coefficients, respectively.
Compare the equations below where the area is assumed to be equal to A = (1.L):
𝐿𝐿
ℎ𝐿𝐿 𝐿𝐿(𝑇𝑇𝑠𝑠 − 𝑇𝑇∞ ) = ∫0 ℎ𝑥𝑥 (𝑇𝑇𝑠𝑠 − 𝑇𝑇∞ )𝑑𝑑𝑑𝑑 (4.32)
Since the temperature difference is constant, it may be taken outside of the integral
and cancelled:
𝐿𝐿
ℎ𝐿𝐿 . 𝐿𝐿 = ∫0 ℎ𝑥𝑥 . 𝑑𝑑𝑑𝑑 (4.33)
𝐿𝐿 𝑘𝑘 𝜌𝜌𝜌𝜌𝜌𝜌 1⁄2
ℎ𝐿𝐿 . 𝐿𝐿 = ∫0 0.332. 𝑥𝑥 � 𝜇𝜇
� . Pr 1⁄3 . 𝑑𝑑𝑑𝑑 (4.34)
Take the constant terms from outside the integral, and divide both sides by k.
𝐿𝐿 𝜌𝜌𝜌𝜌 1⁄2 𝐿𝐿 1 0.5
ℎ𝐿𝐿 . 𝑘𝑘 = 0.332. � 𝜇𝜇 � . Pr 1⁄3 ∫0 �𝑥𝑥� . 𝑑𝑑𝑑𝑑 (4.35)
The left side is defined as the average Nusselt number, NuL. Algebraically rearrange
the right side.
0.332 𝜌𝜌𝜌𝜌 1⁄2 𝜌𝜌𝜌𝜌 1⁄2
Nu𝐿𝐿 = 0.5
. � 𝜇𝜇 � . Pr 1⁄3 . 𝐿𝐿0.5 = 0.664. � 𝜇𝜇 � . Pr 1⁄3 (4.37)
The term in the brackets may be recognized as the Reynolds number, evaluated at
the end of the convective section. Finally,
⁄
Nu𝐿𝐿 = 0.664. 𝑅𝑅𝑅𝑅𝐿𝐿1 2 . Pr 1⁄3 (4.38)
This is our average correlation for laminar flow over a flat plate with constant wall
temperature.
15
For turbulent flow,
𝒉𝒉𝒉𝒉
𝐍𝐍𝐍𝐍 = 𝒌𝒌
= 𝟎𝟎. 𝟎𝟎𝟎𝟎𝟎𝟎𝐑𝐑𝐑𝐑𝟎𝟎.𝟖𝟖
𝑳𝑳 . 𝐏𝐏𝐏𝐏
𝟏𝟏⁄𝟑𝟑
(4.39)
We can develop a turbulent heat transfer correlation in a manner similar to the von
Karman analysis as we did for laminar flow. It is probably easier, having developed
the Reynolds analogy, to follow that course. The local fluid friction factor, Cf,
associated with turbulent flow over a flat plate is given as:
Cf = 0.0592/Rex0.2 (4.40)
The result of the above integration gives the average Nusselt number over the entire
plate:
1⁄3
Nu𝐿𝐿 = 0.037(Re0.8
𝐿𝐿 − 871)Pr (4.43)
Rearrange to get
1⁄3
Nu𝑥𝑥 = 0.0296Re0.8
𝑥𝑥 . Pr (4.44)
Eqs. (4.30) and (4.44) are for isothermal and smooth surfaces. Also, the local friction
and heat transfer coefficients are higher in turbulent flow than they are in laminar flow.
Liquid metals such as mercury have high thermal conductivities, and they
are commonly used in applications that require high heat transfer rates. Liquid
metals have very small Prandtl numbers, for this case the thermal boundary layer
develops much faster than the velocity boundary layer. Then we can assume the
velocity in the thermal boundary layer to be constant at the free stream value and solve
the energy equation. It gives
It is desirable to have a single correlation that applies to all fluids and all Prandtl
numbers, including liquid metals.
16
𝟏𝟏⁄𝟐𝟐
𝒉𝒉𝒙𝒙 𝒙𝒙 𝟎𝟎.𝟑𝟑𝟑𝟑𝟑𝟑𝐏𝐏𝐏𝐏𝟏𝟏⁄𝟑𝟑 𝐑𝐑𝐑𝐑𝑳𝑳
𝑵𝑵𝑵𝑵𝒙𝒙 = = 𝟏𝟏⁄𝟒𝟒 (4.47)
𝒌𝒌 �𝟏𝟏+(𝟎𝟎.𝟎𝟎𝟎𝟎𝟎𝟎𝟎𝟎⁄𝐏𝐏𝐏𝐏)𝟐𝟐⁄𝟑𝟑 �
Flow across cylinders and spheres is frequently encountered in practice. For example,
the tubes in a shell-and-tube heat exchanger involve both internal flow through the
tubes and external flow over the tubes, and both flows must be considered in the
analysis of the heat exchanger.
The characteristic length for a circular cylinder or sphere is taken to be the external
diameter D. Thus, the Reynolds number is defined as Re = ReD = VD/ν where V is the
uniform velocity of the fluid as it approaches the cylinder or sphere. The critical
Reynolds number for flow across a circular cylinder or sphere is about Recr ≅ 2 × 105.
That is, the boundary layer remains laminar for about Re ≲ 2 ×105 and becomes
turbulent for Re ≳ 2 × 105.
Flows across cylinders and spheres, in general, involve flow separation, which is
difficult to handle analytically. Therefore, flow across cylinders and spheres has been
studied experimentally by numerous investigators, and several empirical correlations
have been developed for the heat transfer coefficient.
This relation is quite comprehensive in that it correlates available data well for RePr >
0.2. The fluid properties are evaluated at the film temperature Tf =0.5(T∞ + Ts ), which
is the average of the free-stream and surface temperatures.
𝟏𝟏⁄𝟒𝟒
𝒉𝒉𝒉𝒉 𝝁𝝁
𝐍𝐍𝐍𝐍𝒔𝒔𝒔𝒔𝒔𝒔 = 𝒌𝒌
= 𝟐𝟐 + �𝟎𝟎. 𝟒𝟒𝐑𝐑𝐑𝐑𝟏𝟏⁄𝟐𝟐 + 𝟎𝟎. 𝟎𝟎𝟎𝟎𝐑𝐑𝐑𝐑𝟐𝟐⁄𝟑𝟑 �𝐏𝐏𝐏𝐏𝟎𝟎.𝟒𝟒 � 𝝁𝝁∞ � (4.49)
𝒔𝒔
which is valid for 3.5 ≤ Re ≤ 8 × 104, 0.7 ≤ Pr ≤ 380 and 1.0 ≤ (μ∞/μs) ≤ 3.2. The fluid
properties are evaluated at the free-stream temperature T∞, except for μs, which is
evaluated at the surface temperature Ts.
The average Nusselt number for flow across cylinders can be expressed compactly
as
𝒉𝒉𝒉𝒉
𝐍𝐍𝐍𝐍𝒄𝒄𝒄𝒄𝒄𝒄 = 𝒌𝒌
= 𝑪𝑪𝐑𝐑𝐑𝐑𝒎𝒎 𝐏𝐏𝐏𝐏 𝒏𝒏 (4.50)
1
where, 𝑛𝑛 = 3 and constants C and m are given in the table.
Liquid or gas flow through pipes or ducts is commonly used in heating and cooling
applications and fluid distribution networks. The fluid in such applications is usually
forced to flow by a fan or pump through a flow section.
Most fluids, especially liquids, are transported in circular pipes. Noncircular pipes are
usually used in applications such as the heating and cooling systems of buildings
where the pressure difference is relatively small, the manufacturing and
installation costs are lower, and the available space is limited for ductwork (Fig. 4.8).
Fig. 4.8: Circular pipes can withstand large pressure differences between the inside
and the outside without undergoing any significant distortion, but noncircular pipes
cannot (Cengel and Ghajar, 2015).
For a fixed surface area, the circular tube gives the most heat transfer for the
least pressure drops.
Although the theory of fluid flow is reasonably well understood, theoretical solutions
are obtained only for a few simple cases such as fully developed laminar flow in a
circular pipe. Therefore, we must rely on experimental results and empirical relations
for most fluid flow problems rather than closed-form analytical solutions.
In internal flow, there is no free stream, and thus the fluid velocity in a tube changes
from zero at the surface because of the no-slip condition to a maximum at the tube
centre. Therefore, it is convenient to work with an average or mean velocity Vavg, which
remains constant for incompressible flow when the cross-sectional area of the tube is
constant.
The value of the average (mean) velocity Vavg at some streamwise cross-section
is determined from the requirement that the conservation of mass principle
𝑚𝑚̇ = 𝜌𝜌𝑉𝑉𝑎𝑎𝑎𝑎𝑎𝑎 𝐴𝐴𝑐𝑐 = ∫𝐴𝐴 𝜌𝜌𝜌𝜌(𝑟𝑟)𝑑𝑑𝐴𝐴𝑐𝑐 (4.51)
𝑐𝑐
where 𝑚𝑚̇ is the mass flow rate, ρ is the density, Ac is the cross-sectional area, and u(r)
is the velocity profile.
In fluid flow, it is convenient to work with an average or mean temperature Tm, which
remains constant at a cross section. The mean temperature Tm changes in the flow
direction whenever the fluid is heated or cooled.
̇
𝐸𝐸𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓 = 𝑚𝑚̇𝑐𝑐𝑝𝑝 𝑇𝑇𝑚𝑚 (4.52)
18
Flow in a tube can be laminar or turbulent, depending on the flow conditions. Fluid flow
is streamlined and thus laminar at low velocities but turns turbulent as the velocity is
increased beyond a critical value. Transition from laminar to turbulent flow does not
occur suddenly; rather, it occurs over some range of velocity where the flow fluctuates
between laminar and turbulent flows before it becomes fully turbulent. Most pipe flows
encountered in practice are turbulent. Laminar flow is encountered when highly
viscous fluids such as oils flow in small diameter tubes or narrow passages. Transition
from laminar to turbulent flow depends on the Reynolds number as well as the degree
of disturbance of the flow by surface roughness, pipe vibrations, and the fluctuations
in the flow. The flow in a pipe is laminar for Re < 2300, fully turbulent for Re >10,000,
and transitional in between.
where Vavg is the average flow velocity, D is the diameter of the tube, and ν = μ/ρ is
the kinematic viscosity of the fluid.
For flow through noncircular tubes, the Reynolds number as well as the Nusselt
number, and the friction factor are based on the hydraulic diameter Dh (Fig. 4.9):
Fig. 4.9: The hydraulic diameter Dh = 4 Ac /p is defined such that it reduces to ordinary
diameter for circular tubes. When there is a free surface, such as in open-channel flow,
the wetted perimeter includes only the walls in contact with the fluid.
𝟒𝟒𝑨𝑨𝒄𝒄
𝑫𝑫𝒉𝒉 = (4.54)
𝒑𝒑
In the absence of any work interactions, the conservation of energy equation for the
steady flow of a fluid in a tube can be expressed as (Fig. 4.10).
𝑸𝑸̇ = 𝒎𝒎̇𝒄𝒄𝒑𝒑 (𝑻𝑻𝒆𝒆 − 𝑻𝑻𝒊𝒊 ) (4.55)
where Ti and Te are the mean fluid temperatures at the inlet and exit of the tube,
respectively, and 𝑄𝑄̇ is the rate of heat transfer to or from the fluid.
Fig. 4.10: The heat transfer to a fluid flowing in a tube is equal to the increase in the
energy of the fluid (Cengel and Ghajar, 2015).
19
In the analysis of flow inside pipes, the thermal conditions at the surface can be
approximated to be constant surface temperature (Ts = const) or constant surface heat
flux (qs = const). The constant surface temperature condition is realized when a phase
change process such as boiling, or condensation occurs at the outer surface of a
tube. The constant surface heat flux condition is realized when the tube is subjected
to radiation or electric resistance heating uniformly from all directions. We may have
either Ts = constant or qs = constant at the surface of a tube, but not both.
In the case of qs = constant, the rate of heat transfer can also be expressed as
Fig. 4.11 shows the variation of the tube surface and the mean fluid temperatures
along the tube for the case of constant surface heat flux.
Fig. 4.11: Variation of the tube surface and the mean fluid temperatures along the tube
for the case of constant surface heat flux (Cengel and Ghajar, 2015).
Thermal entrance region is the region of flow over which the thermal boundary layer
develops and reaches the tube centre while thermally fully developed region is the
region beyond the thermal entrance region in which the dimensionless temperature
profile remains unchanged.
20
In the case of Ts = constant, the rate of heat transfer to or from a fluid flowing in a tube
can also be expressed as from Newton's law of cooling:
𝑸𝑸̇ = 𝒒𝒒̇ 𝒔𝒔 𝑨𝑨𝒔𝒔 ∆𝑻𝑻𝒂𝒂𝒂𝒂𝒂𝒂 = 𝒉𝒉𝑨𝑨𝒔𝒔 (𝑻𝑻𝒔𝒔 − 𝑻𝑻𝒎𝒎 )𝒂𝒂𝒂𝒂𝒂𝒂 (𝑾𝑾) (4.60)
where Tb = (Ti +Te )/2 is the bulk mean fluid temperature, which is the
arithmetic average of the mean fluid temperatures at the inlet and the exit of the tube.
It should be noted that the fluid properties in internal flow are usually evaluated at the
bulk mean fluid temperature.
By using arithmetic mean temperature difference, we assume that the mean fluid
temperature varies linearly along the tube, which is hardly ever the case when Ts =
constant. This simple approximation often gives acceptable results, but not
always. Therefore, we need a better way to evaluate ΔTavg.
Fig. 4.12: The variation of the mean fluid temperature along the tube for the case of
constant temperature.
This dimensionless parameter is called the number of transfer units, denoted by NTU,
and is a measure of the effectiveness of the heat transfer systems.
21
𝑻𝑻𝒊𝒊 −𝑻𝑻𝒆𝒆 ∆𝑻𝑻 −∆𝑻𝑻𝒊𝒊
∆𝑻𝑻𝒍𝒍𝒍𝒍 = 𝒍𝒍𝒍𝒍[(𝑻𝑻 = 𝒍𝒍𝒍𝒍(∆𝑻𝑻𝒆𝒆 (4.62)
𝒔𝒔 −𝑻𝑻𝒆𝒆 )⁄(𝑻𝑻𝒔𝒔 −𝑻𝑻𝒊𝒊 )] 𝒆𝒆 ⁄∆𝑻𝑻𝒊𝒊 )
Note that ΔTi = Ts − Ti and ΔTe = Ts −Te are the temperature differences between the
surface and the fluid at the inlet and the exit of the tube, respectively.
This ΔTlm relation appears to be prone to misuse, but it is practically fail-safe, since
using Ti in place of Te and vice versa in the numerator and/or the denominator will,
at most, affect the sign, not the magnitude. Also, it can be used for both heating (Ts >
Ti and Te) and cooling (Ts < Ti and Te) of a fluid in a tube.
A quantity of interest in the analysis of pipe flow is the pressure drop ΔP since it is
directly related to the power requirements of the fan or pump to maintain flow. In
practice, it is convenient to express the pressure loss for all types of fully developed
internal flows (laminar or turbulent flows, circular or noncircular pipes, smooth or rough
surfaces, horizontal or inclined pipes) as (Fig. 4.13):
𝑳𝑳 𝝆𝝆𝑽𝑽𝟐𝟐𝒂𝒂𝒂𝒂𝒂𝒂
𝐏𝐏𝐏𝐏𝐏𝐏𝐏𝐏𝐏𝐏𝐏𝐏𝐏𝐏𝐏𝐏 𝐥𝐥𝐥𝐥𝐥𝐥𝐥𝐥: ∆𝑷𝑷𝑳𝑳 = 𝒇𝒇 𝑫𝑫 (4.63)
𝟐𝟐
where ρV2avg/2 is the dynamic pressure and f is the Darcy friction factor,
𝟖𝟖𝝉𝝉
𝒇𝒇 = 𝝆𝝆𝑽𝑽𝟐𝟐𝒘𝒘 (4.64)
𝒂𝒂𝒂𝒂𝒂𝒂
Fig. 4.13: The relation for pressure loss (and head loss) is one of the most
general relations in fluid mechanics, and it is valid for laminar or turbulent flows,
circular or noncircular tubes, and pipes with smooth or rough surfaces (Cengel and
Ghajar, 2015).
The friction factor for fully developed laminar flow in a circular tube is given as
𝟔𝟔𝟔𝟔𝟔𝟔 𝟔𝟔𝟔𝟔
𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂 𝐭𝐭𝐭𝐭𝐭𝐭𝐭𝐭, 𝐥𝐥𝐥𝐥𝐦𝐦𝐦𝐦𝐦𝐦𝐦𝐦𝐦𝐦: 𝒇𝒇 =
𝝆𝝆𝑽𝑽𝒂𝒂𝒂𝒂𝒂𝒂 𝑫𝑫
=
𝐑𝐑𝐑𝐑
(4.65)
22
Eq. (4.65) shows that in laminar flow, the friction factor is a function of the Reynolds
number only and is independent of the roughness of the pipe surface.
Pressure losses are commonly expressed in terms of the equivalent fluid column
height, called the head loss hL.
∆𝑷𝑷𝑳𝑳 𝑳𝑳 𝑽𝑽𝟐𝟐𝒂𝒂𝒂𝒂𝒂𝒂
𝐇𝐇𝐇𝐇𝐇𝐇𝐇𝐇 𝐥𝐥𝐥𝐥𝐥𝐥𝐥𝐥: 𝒉𝒉𝑳𝑳 = 𝝆𝝆𝝆𝝆
= 𝒇𝒇
𝑫𝑫 𝟐𝟐𝟐𝟐
(4.66)
Two different solutions of Nusselt number are found depending on the physical
situation. For constant heat flux tube flow, the Nusselt number is given by:
𝒉𝒉𝒉𝒉
𝐍𝐍𝐍𝐍 =
𝒌𝒌
= 𝟒𝟒. 𝟑𝟑𝟑𝟑 (4.67)
For fully developed laminar flow in a circular tube subjected to constant surface heat
flux, the Nusselt number is a constant. There is no dependence on the Reynolds or
the Prandtl numbers.
𝒉𝒉𝒉𝒉
𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂𝐂 𝐭𝐭𝐭𝐭𝐭𝐭𝐭𝐭, 𝐥𝐥𝐥𝐥𝐥𝐥𝐥𝐥𝐥𝐥𝐥𝐥𝐥𝐥 (𝑻𝑻𝒔𝒔 = 𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄): 𝐍𝐍𝐍𝐍 = 𝒌𝒌
= 𝟑𝟑. 𝟔𝟔𝟔𝟔 (4.68)
The thermal conductivity k for use in the Nu relations should be evaluated at the bulk
mean fluid temperature.
Nusselt number relations for fully developed laminar flow in noncircular tubes of
various cross sections are tabulated. The Reynolds and Nusselt numbers for flow in
these tubes are based on the hydraulic diameter Dh = 4Ac/p. Once the Nusselt number
is available, the convection heat transfer coefficient is determined from ℎ = 𝑘𝑘Nu⁄𝐷𝐷ℎ .
Determination of the heat transfer coefficient of turbulent tube flow is much more
involved than that for laminar flow. Hence, greater emphasis is usually placed on
empirical relations.
The classic expression for local Nusselt in turbulent pipe flow is due to Colburn, which
is expressed as:
23
When the variation in properties is large due to a large temperature difference, the
following equation can be used
This is called Tate-Sieder equation and here all properties are evaluated at Tb except
μs, which is evaluated at Ts.
Some simple heat transfer devices consist of two concentric tubes and are properly
called double-tube heat exchangers (Fig. 4.14). In such devices, one fluid flows
through the tube while the other flows through the annular space. The governing
differential equations for both flows are identical.
Fig. 4.14: A double-tube heat exchanger that consists of two concentric tubes (Cengel
and Ghajar, 2015).
Consider a concentric annulus of inner diameter Di and outer diameter Do. The
hydraulic diameter of the annulus is
Annular flow is associated with two Nusselt numbers— Nui on the inner tube surface
and Nuo on the outer tube surface—since it may involve heat transfer on both
surfaces. The Nusselt numbers for fully developed laminar flow with one surface
isothermal and the other adiabatic are tabulated.
When Nusselt numbers are known, the convection heat transfer coefficients for the
inner and the outer surfaces are determined from
ℎ𝑖𝑖 𝐷𝐷ℎ ℎ𝑜𝑜 𝐷𝐷ℎ
Nu𝑖𝑖 = and Nu𝑜𝑜 = (4.73)
𝑘𝑘 𝑘𝑘
For fully developed turbulent flow, hi and ho are approximately equal, and the tube
annulus can be treated as a noncircular duct with a hydraulic diameter of Dh = Do − Di.
24
4.10 FREE OR NATURAL CONVECTION
If fluid motion is caused by buoyancy forces within the fluid, it results in what is called
free or natural convection. Buoyancy is due to combined presence of a fluid density
gradient and body force which is proportional to density.
Many familiar heat transfer applications involve natural convection as the primary
mechanism of heat transfer. The motion that results from the continual replacement of
the heated air in the vicinity of the egg by the cooler air nearby is called a natural
convection current, and the heat transfer that is enhanced because of this current is
called natural convection heat transfer.
Generally, natural convection velocities are much smaller than those associated
with forced convection resulting in smaller heat transfer coefficients.
In a gravitational field, there is a net force that pushes upward a light fluid placed in a
heavier fluid. The upward force exerted by a fluid on a body completely or partially
immersed in it is called the buoyancy force. The magnitude of the buoyancy force is
equal to the weight of the fluid displaced by the body. That is,
where ρfluid is the average density of the fluid (not the body), g is the
gravitational acceleration, and Ѵbody is the volume of the portion of the body immersed
in the fluid (for bodies completely immersed in the fluid, it is the total volume of the
body). In the absence of other forces, the net vertical force acting on a body is the
difference between the weight of the body and the buoyancy force. That is,
𝐹𝐹𝑛𝑛𝑛𝑛𝑛𝑛 = 𝑊𝑊 − 𝐹𝐹𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 = 𝜌𝜌𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 𝑔𝑔Ѵ𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 − 𝜌𝜌𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓 𝑔𝑔Ѵ𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 = �𝜌𝜌𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 − 𝜌𝜌𝑓𝑓𝑓𝑓𝑢𝑢𝑖𝑖𝑖𝑖 �𝑔𝑔Ѵ𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 (4.75)
The buoyancy effect has far-reaching implications in life. Without buoyancy, heat
transfer between a hot (or cold) surface and the fluid surrounding it would be by
conduction instead of by natural convection.
25
requires a knowledge of a property that represents the variation of the density of a
fluid with temperature at constant pressure. The property that provides that information
is the volume expansion coefficient β, defined as (Fig. 4.16):
𝟏𝟏 𝝏𝝏𝝏𝝏 𝟏𝟏 𝝏𝝏𝝏𝝏
𝜷𝜷 = 𝝂𝝂 �𝝏𝝏𝝏𝝏� = − 𝝆𝝆 �𝝏𝝏𝝏𝝏� (𝟏𝟏⁄𝐊𝐊) (4.76)
𝑷𝑷 𝑷𝑷
In natural convection studies, the condition of the fluid sufficiently far from the hot or
cold surface and the volume expansion coefficient can be expressed approximately
by replacing differential quantities with differences as
1 ∆𝜌𝜌 1 𝜌𝜌 −𝜌𝜌
𝛽𝛽 ≈ − 𝜌𝜌 ∆𝑇𝑇 = − 𝜌𝜌 𝑇𝑇∞−𝑇𝑇 (𝑎𝑎𝑎𝑎 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 𝑃𝑃) (4.77)
∞
Or
𝝆𝝆∞ − 𝝆𝝆 = 𝝆𝝆𝝆𝝆(𝑻𝑻 − 𝑻𝑻∞ ) (𝒂𝒂𝒂𝒂 𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄𝒄 𝑷𝑷) (4.78)
where ρ∞ is the density and T∞ is the temperature of the quiescent fluid away from the
surface.
It can easily be shown that the volume expansion coefficient β of an ideal gas
(P = ρRT) at a temperature T is equivalent to the inverse of the temperature:
𝟏𝟏
𝜷𝜷𝒊𝒊𝒊𝒊𝒊𝒊𝒊𝒊𝒊𝒊 𝒈𝒈𝒈𝒈𝒈𝒈 = 𝑻𝑻 (4.79)
The larger the temperature difference between the fluid adjacent to a hot (or cold)
surface and the fluid away from it, the larger the buoyancy force and the stronger the
natural convection currents, and thus the higher the heat transfer rate.
In natural convection, no blowers are used, and therefore the flow rate cannot be
controlled externally. The flow rate in this case is established by the dynamic balance
of buoyancy and friction.
For solids and liquids, β can be found from property tables. Values of β are found for
several engineering fluids in tables given in handbooks and textbooks.
26
Coefficient of volumetric expansion is the thermodynamic property which describes
the change in density leading to buoyancy.
Because V∞ is always zero, the Reynolds number, [ρV∞D]/μ, is also zero and is no
longer suitable to describe the flow in the system. Instead, we introduce a new
parameter for natural convection, the Grashof Number. Here we will be most
concerned with flow across a vertical surface, so that we use the vertical distance, Lc,
as the characteristic length.
The Grashof number provides the main criterion in determining whether the fluid flow
is laminar or turbulent in natural convection. For vertical plates, the critical Grashof
number is observed to be about 109.
When a surface is subjected to external flow, the problem involves both natural
and forced convection. The relative importance of each mode of heat transfer is
determined by the value of the coefficient Gr/Re2:
• Free convection dominates and the forced convection effects are negligible if
Gr/Re2 >> 1.
• Both effects are significant and must be considered if Gr/Re2 ≈ 1(mixed convection).
Natural convection heat transfer on a surface depends on the geometry of the surface
as well as its orientation, the variation of temperature on the surface and the
thermophysical properties of the fluid involved.
27
Except for some simple cases, heat transfer relations in natural convection are based
on experimental studies. Quite often from these experiments, it has been found that
the exponent on the Grashof and Prandtl numbers are equal so that the general
correlations may be written in the form:
This leads to the introduction of the new, dimensionless parameter, the Rayleigh
number, Ra:
Ra𝐿𝐿 = Gr𝐿𝐿 Pr
or
𝒉𝒉𝑳𝑳𝒄𝒄
𝐍𝐍𝐍𝐍 = = 𝑪𝑪(𝐆𝐆𝐆𝐆𝑳𝑳 𝐏𝐏𝐏𝐏)𝒏𝒏 = 𝑪𝑪𝐑𝐑𝐑𝐑𝐧𝐧𝐋𝐋 (4.83)
𝒌𝒌
The constants C and n depend on the geometry of the surface and the flow regime,
which is characterized by the range of the Rayleigh number.
The value of n is usually 1/4 for laminar flow and 1/3 for turbulent flow. All fluid
properties are to be evaluated at the film temperature Tf = (Ts + T∞)/2. In all cases, the
fluid properties should be evaluated at film temperature.
When the average Nusselt number and thus the average convection coefficient is
known, the rate of heat transfer by natural convection from a solid surface at a uniform
temperature Ts to the surrounding fluid is expressed by Newton’s law of cooling as
where, As is the heat transfer surface area, and h is the average heat
transfer coefficient on the surface.
1. Vertical plates
For vertical plates both at constant temperature and constant flux, Churchill and Chu
equation is used for both laminar and turbulent flow (entire range of Rayleigh number)
respectively:
𝟏𝟏⁄𝟔𝟔
𝟐𝟐
𝟎𝟎.𝟑𝟑𝟑𝟑𝟑𝟑𝐑𝐑𝐑𝐑𝑳𝑳
𝐍𝐍𝐍𝐍 = �𝟎𝟎. 𝟖𝟖𝟖𝟖𝟖𝟖 + 𝟖𝟖⁄𝟐𝟐𝟐𝟐 � (4.85)
�𝟏𝟏+(𝟎𝟎.𝟒𝟒𝟒𝟒𝟒𝟒⁄𝐏𝐏𝐏𝐏)𝟗𝟗⁄𝟏𝟏𝟏𝟏 �
This relation is most accurate in the range of 10−1 < RaL < 109.
28
2. Inclined plates
For an inclined hot plate that makes an angle θ from the vertical, the net force 𝐹𝐹 =
𝑔𝑔(𝜌𝜌∞ − 𝜌𝜌) (the difference between the buoyancy and gravity) acting on a unit volume
of the fluid in the boundary layer is always in the vertical direction. In the case of an
inclined plate, this force can be resolved into two components: 𝐹𝐹𝑥𝑥 = 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹 parallel to
the plate that drives the flow along the plate, and 𝐹𝐹𝑦𝑦 = 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹 normal to the plate.
Fig. 4.17: Natural convection flows on the upper and lower surfaces of an inclined hot
plate.
In a hot plate in a cooler environment for the lower surface of a hot plate, the
convection currents are weaker, and the rate of heat transfer is lower relative to the
vertical plate case.
On the upper surface of a hot plate, the thickness of the boundary layer and thus the
resistance to heat transfer decreases, and the rate of heat transfer increases relative
to the vertical orientation. In the case of a cold plate in a warmer environment, the
opposite occurs.
When the boundary layer remains intact (the lower surface of a hot plate or the upper
surface of a cold plate), the Nusselt number can be determined from the vertical plate
relations provided that g in the Rayleigh number relation is replaced by 𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔 for θ <
60°.
Generally, for an inclined plate with an angle θ to the vertical, use of vertical plate
equations for the upper surface of a cold plate and the lower surface of a hot plate.
3. Horizontal plates
For a hot surface in a cooler environment, the net force acts upward, forcing the heated
fluid to rise (see Fig. 4.18).
29
Fig. 4.18: Natural convection flows on the upper and lower surfaces of a horizontal
hot plate (Cengel and Ghajar, 2015).
If the hot surface is facing upward, the heated fluid rises freely, inducing strong natural
convection currents and thus effective heat transfer. Therefore, depending on the Ra
range, use either
⁄4
Nu = 0.59Ra1𝐿𝐿 (for 104 to 107 ) (4.86)
or
⁄3
Nu = 0.15Ra1𝐿𝐿 (for 107 to 1011 ) (4.87)
But if the hot surface is facing downward, the plate blocks the heated fluid that tends
to rise, impeding heat transfer. Then,
⁄4
Nu = 0.27Ra1𝐿𝐿 (for 105 to 1011 ) (4.88)
The opposite is true for a cold plate in a warmer environment since the net force
(weight minus buoyancy force) in this case acts downward, and the cooled fluid near
the plate tends to descend.
Note that 𝐿𝐿𝑐𝑐 = 𝑎𝑎⁄4 for a horizontal square surface of length a and 𝐿𝐿𝑐𝑐 = 𝐷𝐷⁄4 for a
horizontal circular surface of diameter D.
4. Vertical cylinders
30
35𝐿𝐿
𝐷𝐷 ≥ 1⁄4 (4.90)
Gr𝐿𝐿
𝐷𝐷 1⁄4 𝐷𝐷
NuD = 0.6 �RaD 𝐿𝐿 � when RaD 𝐿𝐿 ≥ 101⁄4 ( 4.91)
𝐷𝐷 0.16 𝐷𝐷
NuD = 1.37 �RaD 𝐿𝐿 � when 0.05 ≤ RaD 𝐿𝐿 ≤ 101⁄4 (4.92)
𝐷𝐷 0.05 𝐷𝐷
NuD = 0.93 �RaD 𝐿𝐿 � when RaD 𝐿𝐿 ≤ 0.05 (4.93)
The boundary layer over a hot horizontal cylinder starts to develop at the bottom,
increasing in thickness along the circumference, and forming a rising plume at the top
as shown in Fig. 4.20.
Fig. 4.20: Natural convection flow over a horizontal hot cylinder (Cengel and Ghajar,
2015).
Therefore, the local Nusselt number is highest at the bottom, and lowest at the top of
the cylinder when the boundary layer flow remains laminar. The opposite is true in the
case of a cold horizontal cylinder in a warmer medium, and the boundary layer in this
case starts to develop at the top of the cylinder and ending with a descending plume
at the bottom.
The average Nusselt number over the entire surface can be determined
from Churchill and Chu equation for an isothermal horizontal cylinder
𝟏𝟏⁄𝟔𝟔
𝟐𝟐
𝟎𝟎.𝟑𝟑𝟑𝟑𝟑𝟑𝐑𝐑𝐑𝐑𝑳𝑳
𝐍𝐍𝐍𝐍 = �𝟎𝟎. 𝟔𝟔 + 𝟖𝟖⁄𝟐𝟐𝟐𝟐 � (4.94)
�𝟏𝟏+(𝟎𝟎.𝟓𝟓𝟓𝟓𝟓𝟓⁄𝐏𝐏𝐏𝐏)𝟗𝟗⁄𝟏𝟏𝟏𝟏 �
The empirical correlations for average Nusselt number for natural convection over
surfaces are given in Table 9-1 of the prescribed textbook.
31
4.11 WORKED EXAMPLES AND REVISION QUESTIONS
Example 4.1
During air cooling of potatoes, the heat transfer coefficient for combined convection,
radiation, and evaporation is determined experimentally to be as shown:
Solution:
Assumptions:
1. Steady operating conditions exist.
2. Potato is spherical in shape.
3. Convection heat transfer coefficient is constant over the entire surface.
Analysis:
32
where the heat transfer coefficient is obtained from the table at 1 m/s velocity. The
initial value of the temperature gradient at the potato surface is:
Example 4.2
An average man has a body surface area of 1.8 m2 and a skin temperature of 33° C.
The convection heat transfer coefficient for a clothed person walking in still air is
expressed as h = 8.6 V0.53 for 0.5 < V < 2 m/s, where V is the walking velocity in m/s.
Assuming the average surface temperature of the clothed person to be 30° C,
determine the rate of heat loss from an average man walking in still air at 7° C by
convection at a walking velocity of (a) 0.5 m/s, (b) 1.0 m/s, (c) 1.5 m/s, and (d) 2.0 m/s.
Solution
Assumptions:
Analysis:
The convection heat transfer coefficients and the rate of heat losses at different
walking velocities are:
33
Example 4.3
During air cooling of oranges, grapefruit, and tangelos, the heat transfer coefficient for
combined convection, radiation, and evaporation for air velocities of
0.11 < V < 0.33 m/s is determined experimentally and is expressed as
h = 5.05 kairRe1/3/D, where the diameter D is the characteristic length. Oranges are
cooled by refrigerated air at 3° C and 1 atm at a velocity of 0.3 m/s. Determine (a) the
initial rate of heat transfer from a 7-cm-diameter orange initially at 15° C with a
thermal conductivity of 0.7 W/m⋅K, (b) the value of the initial temperature gradient
inside the orange at the surface, and (c) the value of the Nusselt number.
Solution
Assumptions:
The thermal conductivity and the kinematic viscosity of air at the film temperature of
(Ts+T∞)/2 = (15+5)/2 = 10°C are (Table A-15):
Analysis:
(a) The Reynolds number, the heat transfer coefficient, and the initial rate of heat
transfer from an orange are
34
(b) The temperature gradient at the orange surface is determined from
The convection heat transfer coefficient for a clothed person standing in moving air is
expressed as h = 14.8 V0.69 for 0.15 < V < 1.5 m/s, where V is the air velocity. For a
person with a body surface area of 1.7 m2 and an average surface temperature of 29°
C, determine the rate of heat loss from the person in windy air at 10° C by convection
for air velocities of (a) 0.5 m/s, (b) 1.0 m/s, and (c) 1.5 m/s.
The upper surface of a 50-cm-thick solid plate (k = 237 W/m⋅K) is being cooled by
water with temperature of 20° C. The upper and lower surfaces of the solid plate
maintained at constant temperatures of 60° C and 120° C, respectively. Determine the
water convection heat transfer coefficient and the water temperature gradient at the
upper plate surface.
Question 4.3
What fluid property is responsible for the development of the velocity boundary layer?
For what kinds of fluids will there be no velocity boundary layer on a flat plate?
Question 4.4
What is the physical significance of the Prandtl number? Does the value of the Prandtl
number depend on the type of flow or the flow geometry? Does the Prandtl number of
air change with pressure? Does it change with temperature?
35
Question 4.5
Will a thermal boundary layer develop in flow over a surface even if both the fluid and
the surface are at the same temperature?
Question 4.6
What is the physical significance of the Reynolds number? How is it defined for
external flow over a plate of length L?
Water at 43.3° C flows over a large plate at a velocity of 30.0 cm/s. The plate is 1.0-m
long (in the flow direction), and its surface is maintained at a uniform temperature of
10.0° C. Calculate the steady rate of heat transfer per unit width of the plate.
Solution
Assumptions:
ρ = 996.6 kg/m3
μ = 0.854 x 10-3 kg/ms
k = 0.610 W/m.ºC
Pr = 5.85
which is smaller than the critical Reynolds number. Thus, we have laminar flow for the
entire plate.
Solution
Assumptions:
Since Rex < 5 x 105, the flow is laminar. Using the proper relation for Nusselt number,
the local heat transfer coefficient at 1 m from the leading edge of the flat plate is
37
Since Rex < 5 x 105, the flow is laminar. Using the proper relation for Nusselt number,
the average heat transfer coefficient of the entire flat plate is
(c) The total heat flux transfer to the flat plate on the upper and lower surfaces is
The forming section of a plastics plant puts out a continuous sheet of plastic that is
1.2-m wide and 2-mm thick at a rate of 15 m/min. The temperature of the plastic sheet
is 90° C when it is exposed to the surrounding air, and the sheet is subjected to airflow
at 30° C at a velocity of 3 m/s on both sides along its surfaces normal to the direction
of motion of the sheet. The width of the air-cooling section is such that a fixed point on
the plastic sheet passes through that section in 2 s. Determine the rate of heat transfer
from the plastic sheet to the air.
Solution
Assumptions:
38
which is less than the critical Reynolds number. Thus, the flow is laminar. Using the
proper relation in laminar flow for Nusselt number, the average heat transfer coefficient
and the heat transfer rate are determined to be:
The top surface of the passenger car of a train moving at a velocity of 95 km/h is 2.8-
m wide and 8-m long. The top surface is absorbing solar radiation at a rate of 380
W/m2, and the temperature of the ambient air is 30° C. Assuming the roof of the car to
be perfectly insulated and the radiation heat exchange with the surroundings to be
small relative to convection, determine the equilibrium temperature of the top
surface of the car. Ans.: 37.5° C
During a cold winter day, wind at 42 km/h is blowing parallel to a 6-m-high and 10-m-
long wall of a house. If the air outside is at 5° C and the surface temperature of the
wall is 12° C, determine the rate of heat loss from that wall by convection. What would
your answer be if the wind velocity was doubled? Ans.: 10.8 kW, 19.4 kW
39
4.11.5 EXAMPLES ON FORCED INTERNAL CONVECTION
Solution
Assumptions:
The properties of air at the anticipated average temperature of 30°C are (Table A-15)
The logarithmic mean temperature difference and the rate of heat transfer are:
40
Example 4.8 (8-26 in prescribed textbook)
Cooling water available at 10° C is used to condense steam at 30° C in the condenser
of a power plant at a rate of 0.15 kg/s by circulating the cooling water through a bank
of 5-m-long, 1.2-cm-internal-diameter thin copper tubes. Water enters the tubes at a
mean velocity of 4 m/s and leaves at a temperature of 24° C. The tubes are nearly
isothermal at 30° C. Determine the average heat transfer coefficient between
the water, the tubes, and the number of tubes needed to achieve the indicated heat
transfer rate in the condenser.
Solution
Assumptions:
The mass flow rate of water and the surface area are
The logarithmic mean temperature difference and the surface area are:
41
The total rate of heat transfer is determined from:
Combustion gases passing through a 5-cm-internal diameter circular tube are used to
vaporize wastewater at atmospheric pressure. Hot gases enter the tube at 115 kPa
and 250° C at a mean velocity of 5 m/s and leave at 150° C. If the average heat
transfer coefficient is 120 W/m2⋅K and the inner surface temperature of the tube is
110° C, determine (a) the tube length and (b) the rate of evaporation of water. Use
air properties for the combustion gases.
Solution
Assumptions:
Also, the heat of vaporization of water at 1 atm or 100°C is hfg = 2431 kJ/kg Table A-
9)
The density of air at the inlet and the mass flow rate of exhaust gases are:
42
The rate of heat transfer is:
The logarithmic mean temperature difference and the surface area are:
Determine the convection heat transfer coefficient for the flow of (a) air and (b) water
at a velocity of 5 m/s in an 8-cm-diameter and 10-m-long tube when the tube is
subjected to uniform heat flux from all surfaces. Use fluid properties at 25° C.
Solution
Assumptions:
which is greater than 10,000. Therefore, the flow is turbulent and the entry lengths in
this case are roughly.
which is much shorter than the total length of the tube. Therefore, we can assume fully
developed turbulent flow in the entire duct, and determine the Nusselt number from
44
Example 4.11 (8-125 in prescribed textbook)
Liquid mercury is flowing at 0.6 kg/s through a 5-cm-diameter tube with inlet and outlet
mean temperatures of 100° C and 200° C, respectively. The tube surface
temperature is kept constant at 250° C. Determine the tube length using (a) the
appropriate Nusselt number relation for liquid metals and (b) the Dittus-Boelter
equation. (c) Compare the results of (a) and (b).
Solution
Assumptions:
(a) The flow is turbulent and the appropriate Nusselt number relation for liquid metals
is
ℎ = 1758 𝑊𝑊 ⁄𝑚𝑚2 . 𝐾𝐾
45
Hence, length of the tube is:
(c) Comparing the results calculated for (a) and (b) showed that when using the Dittus-
Boelter equation, the tube length is underestimated by approximately 36%.
In the effort to find the best way to cool a smooth, thin-walled copper tube, an engineer
decided to flow air either through the tube or across the outer tube surface. The tube
has a diameter of 5 cm, and the surface temperature is held constant. Determine (a)
the convection heat transfer coefficient when air is flowing through its inside at 25 m/s
with a bulk mean temperature of 50° C and (b) the convection heat transfer coefficient
when air is flowing across its outer surface at 25 m/s with a film temperature of 50° C.
Solution
Assumptions:
The properties of air at 50°C: k = 0.02735 W/m∙K, ν = 1.798 x 10−5 m2/s, and Pr =
0.7228 (Table A-15).
(a) The flow inside the tube is turbulent (Re > 10,000), and using the Dittus-Boelter
equation, we have
46
(b) Using the Zukauskas (1972) equation from Table 7-1, we have
Solution
Assumptions:
The properties of water at the bulk mean fluid temperature of (15+35)/2=25ºC are:
(Table A-9)
The properties of water at 25°C are (Table A-9): ρ = 997 kg/m3, μ = 0.891 x 10-3 kg/ms,
47
which is greater than 10,000. Therefore, we have turbulent flow. Assuming fully
developed flow in the entire tube, the Nusselt number is determined from:
Considering that there are 5 tubes, the logarithmic mean temperature difference, the
surface area and the length of the tubes are determined as follows:
Repeat Prob. 8–26 for steam condensing at a rate of 0.60 kg/s. Ans.: 12.1 W/m2.K;
55.
Repeat Prob. 8–24 for a heat transfer coefficient of 40 W/m2⋅K. Ans.: 153 cm; 1.23
kg/h.
48
4.11.7 EXAMPLES ON FREE OR NATURAL CONVECTION
Show that the volume expansion coefficient of an ideal gas is β = 1/T, where T is the
absolute temperature.
Solution
Or
Thus,
Using its definition and the values listed in Table A–9, determine the volume
expression coefficient of saturated liquid water at 70° C. Compare the result with the
value tabulated in Table A–9.
Solution
Assumption:
The properties of saturated liquid water are listed in the following table:
49
The volume expansion coefficient is defined as
A 10-cm × 10-cm plate has a constant surface temperature of 150° C. Determine the
Grashof number if the chip is placed in the following fluids: air (1 atm, 30° C), liquid
water (30° C), engine oil (10° C). Discuss how the Grashof number affects the natural
convection flow.
Solution
Assumption:
The properties of air, liquid water, and engine oil are listed in the following table:
50
For air,
The Grashof number for liquid water and engine oil are calculated like the calculation
done for air above,
Solution
Assumption:
The properties of air at Tf = (Ts + T∞)/2 = 20°C are k = 0.02514 W/m∙K, n = 1.516 ×
10−5 m2/s, Pr = 0.7309 (from Table A-15). Also, β = 1/Tf = 0.003413 K-1.
The Rayleigh number (Lc = D, i.e., for horizontal cylinder, the characteristic length is
its diameter) is
51
Then, the heat transfer coefficient is,
A can of engine oil with a length of 150 mm and a diameter of 100 mm is placed
vertically in the trunk of a car. On a hot summer day, the temperature in the trunk is
43° C. If the surface temperature of the can is 17° C, determine heat transfer rate from
the can surface. Neglect the heat transfer from the ends of the can.
Solution
Assumption:
Properties The properties of air at Tf = (Ts + T∞)/2 = 30°C are k = 0.02588 W/m∙K, ν =
1.608 × 10−5 m2/s, Pr = 0.7282 (from Table A-15). Also, β = 1/Tf = 0.0033 K-1.
Analysis:
The Rayleigh number (Lc = L, i.e. For vertical cylinder, the characteristic length is its
length) is,
Then,
52
⁄
Since 𝐷𝐷 ≥ 35 𝐿𝐿⁄𝐺𝐺𝐺𝐺𝐿𝐿1 4 is satisfied, we can treat this vertical cylinder as a vertical plate,
and the Nusselt may be calculated with
Consider a vertical plate with length L, placed in quiescent air. If the film temperature
is 20° C and the average Nusselt number in natural convection is of the form Nu =
𝐶𝐶Ra𝑛𝑛𝐿𝐿 , show that the average heat transfer coefficient can be expressed as
ℎ = 1.51(∆𝑇𝑇/𝐿𝐿)1⁄4 104 < 𝑅𝑅𝑅𝑅𝐿𝐿 < 109
ℎ = 1.19∆𝑇𝑇 1⁄3 1010 < 𝑅𝑅𝑅𝑅𝐿𝐿 < 1013
Solution
The properties of air at Tf = 20°C are: k = 0.02514 W/m∙K, ν = 1.516 × 10−5 m2/s, Pr =
0.7309 (from Table A-15). Also, β = 1/Tf = 0.003413 K-1.
53
Then,
Then,
A plate (0.5 m × 0.5 m) is inclined at an angle of 30°. The top surface of the plate is
well insulated. Estimate the rate of heat loss from the plate when the bottom surface
is maintained at 60°C and the surrounding atmospheric quiescent air is at 0° C.
Solution
Assumption:
Properties The properties of air at Tf = (Ts + T∞)/2 = 30°C are: k = 0.02588 W/m∙K, ν =
1.608 × 10−5 m2/s, Pr = 0.7282 (from Table A-15). Also, β = 1/Tf = 0.0033 K-1.
54
The Rayleigh number (Lc = L) is
The Nusselt number is calculated using the correlation for vertical plate,
Solution
Assumptions:
The properties of air at 1 atm and the film temperature of (Ts+T¥)/2 = (40+10)/2 = 25°C
are (Table A-15)
55
ν = 1.562 x 10-5 m2/s, k = 0.02551 W/m.ºC, Pr = 0.7296
(a) When the stack is exposed to 10 m/s winds, the heat transfer will be by forced
convection. We have flow of air over a cylinder and the heat transfer rate is determined
as follows:
(b) Without wind the heat transfer will be by natural convection. The characteristic
length in this case is the outer diameter of the cylinder, Lc = D = 0.12 m. Then,
Consider a flat-plate solar collector placed horizontally on the flat roof of a house. The
collector is 1.5 m wide and 4.5 m long, and the average temperature of the
exposed surface of the collector is 42° C. Determine the rate of heat loss from the
collector by natural convection during a calm day when the ambient air temperature is
56
8 °C. Also, determine the heat loss by radiation by taking the emissivity of the
collector surface to be 0.85 and the effective sky temperature to be – 15 °C.
Solution
Assumptions:
The properties of air at 1 atm and the film temperature of (Ts+T∞)/2 = (42+8)/2 = 25°C
are (Table A-15):
Then,
and
57
Heat transfer rate by radiation is:
Question 4.13
The density of liquid water can be correlated as ρ(T) = 1000 − 0.0736T − 0.00355
T2 where ρ and T are in kg/m3 and °C, respectively. Determine the volume
expansion coefficient at 70° C. Compare the result with the value tabulated in Table
A–9. Ans.: 5.84 x 10-4 K-1.
Flue gases from an incinerator are released to atmosphere using a stack that is 0.6 m
in diameter and 10.0 m high. The outer surface of the stack is at 40° C and the
surrounding air is at 10° C. Determine the rate of heat transfer from the stack,
assuming (a) there is no wind and (b) the stack is exposed to 20 km/h winds. Ans.:
(a) 2070 W; (b) 11390 W
Consider an L × L horizontal plate that is placed in quiescent air with the hot surface
facing up. If the film temperature is 20° C and the average Nusselt number in
natural convection is of the form Nu = 𝐶𝐶Ra𝑛𝑛𝐿𝐿 , show that the average heat transfer
coefficient can be expressed as
58