You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/258697089

Handbook of Burner Technology for Industrial Furnaces

Chapter · January 2009

CITATIONS READS

0 7,596

1 author:

Ulrich Renz
RWTH Aachen University
232 PUBLICATIONS   1,670 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

I retired from Aachen University nearly 15 years ago. View project

All content following this page was uploaded by Ulrich Renz on 02 April 2016.

The user has requested enhancement of the downloaded file.


III

Joachim G. Wünning, Ambrogio Milani (ed.)

Handbook of
Burner Technology for
Industrial Furnaces
Fundamentals | Burner | Applications

2nd Edition

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

3. Heat transfer 31

3. Heat transfer
Ulrich Renz

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

32 3. Heat transfer

3. Heat transfer
Ulrich Renz

3.1 Mechanisms of heat transfer


The methods of heat transfer in gaseous, liquid or solid mediums or between two mediums can
be classified into:
• Radiation
• Conduction
• Convection.
The physical principles of these three methods will only be outlined in this introduction.
If two bodies of different temperatures are insulated in an evacuated container, we can expect
temperature equilibrium after a certain length of time. This process of temperature equilibrium is
based on heat transfer through electromagnetic waves – radiation.
Thermodynamics show that an ideal body, the so-called “black” body radiates heat, which is pro-
portional to the fourth power of its absolute temperature.
When radiation between technical surfaces takes place, the deviation from the emissivity of an
ideal “black” body must be taken into account. This is done by introducing the deviation factor F¡
and an additional factor FA, which represents deviations that occur having in mind the fact that the
emitted radiation from one body is not necessarily fully absorbed by the other due to their mutual
orientation in space.
Determining the values of these factors can be very complicated and will be discussed later on
in detail. Much more elaborate are the processes of radiation that involve a gaseous atmosphere
between the two bodies.
If there is a difference in temperature within a body, then we expect heat to be transferred from
one area of higher temperature to the area where the temperature is lower. The heat flow per area
is proportional to the temperature gradient. The proportionality constant is characteristic of the
material and is called the thermal conductivity h [W/mK]:

(3.1)

The minus sign indicates that the heat flows in the direction of the temperature gradient.

Figure 3.1: Heat transfer between two bodies


through radiation

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

3. Heat transfer 33

In order to determine the heat flow quan- Table 3.1: Thermal conductivity of selected materials
titatively, the thermal conductivity must be at 0 °C
known. In gases the thermal conductivity
Thermal
is determined by the collisions of the gas
Materials conductivity
molecules, where the kinetic energy of a
h (W/mk)
gas molecule is presumed proportional
to its temperature. Moving from one area Metals
of higher temperature to an area of lower Silver 410
temperature, the molecule loses part of its
energy and momentum through collisions. If Copper 385
we assume the gas to be a continuum then Carbon-enriched steel 43
we can consider this process a process of
Chromium-Nickel-enriched
heat conduction in gases. 16
steel 3
For simple gas molecules and moderate Non-metallic solids
temperatures the thermal conductivity can
be calculated on the basis of the kinetic Quartz 2.4
theory. On the other hand, liquids and solids Sand-rock 1.8
require more precise assumptions than the
above theoretical premises so that in gen- Window glass 0.8
eral the values of the thermal conductivity Liquids
are determined experimentally.
Mercury 8.21
Table 3.1 lists a few thermal conductivity
Water 0.56
constants to show the order of their values
for different mediums. Cooling agent R 12 0.07
It is generally known that a hot object cools Gases
faster when it is being blown upon as com- Hydrogen 0.175
pared to it being left to freely cool down in
open air. The heat conductivity mechanism Air 0.024
on the part of the air side of the system is Water vapor 0.021
additionally reinforced with the mechanism
of heat transfer through the macroscopic air
flow along the body (Figure 3.2). The intensity of this so-called convective heat transfer is pro-
portional to the temperature difference between the body surface area and the cooling medium.
The proportionality constant – the convection heat-transfer coefficient _ – depends on many pa-
rameters such as the flow velocity, the medium type, etc. Even so, the convection heat-transfer
coefficient can be determined theoretically in the simplest cases using the principles of flow dy-
namics. Yet, in many real world situations, it can only be determined experimentally.

Figure 3.2: Convective heat flux

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

34 3. Heat transfer

Table 3.2: Order of magnitude of the convection heat-transfer coefficient

Convection heat-transfer coefficient _ [W/m2K]


Free convection Gases 3 – 20
Water 100 – 600
Forced convection Gases 10 – 100
Water 500 – 10,000
Boiling water 2,000 – 25,000
Water vapor condensation 5,000 – 10,000

The order of magnitude of the convection heat-transfer coefficient is listed in Table 3.2.
In industrial furnaces these three basic mechanisms of heat transfer happen simultaneously. The
heat transmission inside the load occurs by heat conduction. Natural convection and radiation
cool the outer furnace wall. The heat transfer from the hot inside furnace walls to the load takes
place by radiation and convection, whereas radiation is clearly dominant at higher furnace temper-
atures. At lower furnace temperatures or low emissivity of the load, heat transfer can be increased
by a targeted flow configuration to increase forced convection. Very high convective heat transfer
coefficients can be achieved with impingement systems (see Figure 3.3). Depending on the nozzle
configuration, convective heat transfer coefficients of well over 100 W/m2K are possible. Impinge-
ment systems are used for heating and defined cooling.

3.2 Heat radiation


3.2.1 Electromagnetic spectrum
Conduction and convection are heat transfer mechanisms where energy is transported through
molecular processes, i.e. macroscopic movement of fluids, whereas heat radiation does not re-
quire any medium since it is based on electromagnetic processes. The intensity and the type of
radiation, emitted from a gaseous, liquid or solid body depend on the surface properties of the
body and its temperature, yet they are independent of its surroundings. If not only the emitted
radiation is to be considered, but also the heat exchange between the body and its surroundings,
then the type, temperature and geometrical orientation in space of the surrounding bodies must be

Figure 3.3: Forced convection through impinging jets

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

3. Heat transfer 35

taken into account. In spite of the fact that in most heat transfer problems energy is simultaneously
transported through conduction and/or convection and radiation, we will deal henceforth, as far as
possible, with radiation only.
Figure 3.4 shows the electromagnetic spectrum and the wavelengths at which heat radiation takes
place.

Figure 3.4: Electromagnetic spectrum

3.2.2 Stefan-Boltzmann’s law


Every body radiates over a wide spectrum of wavelengths, whose intensity depends on the abso-
lute temperature of the body.
The net radiation emitted and the absolute temperature of the body are related by a simple relation-
ship for a so-called “black body”, which in turn is defined as a body that absorbs all the radiation
falling upon it. The Stefan-Boltzmann law, valid for black bodies is

(3.2)

where b indicates the black body.


The Stefan-Boltzmann constant has the value m = 5.67 · 10-8 [W/m2K4].
In practice, an alternative form of the Stefan-Boltzmann law is used:

(3.3)

3.2.3 Planck’s distribution law


Planck derived from the theory of quantum mechanics the relationship for the distribution of the
radiation intensity over the wavelength of a black body.

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

36 3. Heat transfer

Planck’s distribution law for monochrome or spectral radiation flux in a narrow wavelength band
d,h is defined as follows:

(3.4)

C1 = 3.741 · 10-16 Wm2

C2 = 1.439 · 10-2 mK

The diagram in Figure 3.5 shows Planck’s


distribution law for selected temperatures
and radiation wavelengths.
As can be seen in Figure 3.5, the emitted
radiation energy increases with increasing
temperature and the corresponding maxi-
mum shifts to smaller wavelengths as the
temperature increases.
The derivative of equation (3.4) gives the
position of the maxima of the curve. This
is the so-called Wien’s law of displace-
ment:

hmax T = 2,898 [+mK] (3.5)

By integrating the spectral emissive pow-


er over the entire wavelength band we get
the total emissive power of the body and a
relationship between Planck’s distribution
law and Stefan Boltzmann law. Figure 3.5: Spectral radiative heat flux of a black body

3.2.4 Reflection, absorption, trans-


mission
When radiation of a specific wavelength
falls upon the surface of a (non-black)
body, then this radiation is either reflect-
ed off the surface, absorbed by the body
or transmitted through it, whereas in the
case of reflection, it is distinguished be-
tween fully reflective (angle of incidence is
equal to the angle of reflection) or diffuse,
in which case the radiation is equally dis-
tributed in all directions (see Figure 3.6).
It is:
l(h) + _(h) + o(h) = 1 (3.6)

with:
l(h) – spectral reflectivity Figure 3.6: Contributions to radiative transport

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

3. Heat transfer 37

_(h) – spectral absorptivity


o(h) – spectral transmissivity

We can distinguish between the following special cases:


• Since most solids are opaque to radiation and hence the incident radiation is partly reflected
and partly absorbed in thin layers of a few mm (electric conductors) up to 2 mm (electric in-
sulators)

l(h) + _(h) = 1 (3.7)

• Solids, which fully absorb all radiation have already been defined as “black bodies”
3
_(h) = _= 1 (3.8)

• “Grey” bodies, on the other hand, are bodies with radiation properties that are independent of
the wavelength (_(h) = _, l(h) = l, o(h) = o) and which radiate in all directions (diffuse radiation)

l+_+o=1 (3.9)

• Gases normally do not reflect radiation, hence

_(h) + o(h) = 1 (3.10)

3.2.5 Kirchhoff’s law


Kirchhoff’s law states the relationship between the absorbed and the emitted radiation of a body.
The common notation for the Kirchhoff’s law is:

_=¡ (3.11)

Strictly speaking, Kirchhoff’s law is valid only for monochromatic radiation, i.e.

_(h) = ¡(h) (3.12)

Therefore, Kirchhoff’s law, defined for the total emissivity ¡ and total absorptivity _, may not always
be valid since the incident and emitted radiation are not equally dependent on the temperature and
spectral distribution.
Two special cases should be mentioned though, for which Kirchhoff’s law for the total radiation
properties remains valid.

• The radiating body is a black or grey body whose temperature is equal to that of the investi-
gated body, Trad = Tbody

• The surfaces of the body are grey, i.e. their absorptivity is independent of the wavelength.

Especially the second case is of importance since for many practical purposes this holds true.
If a high level of accuracy is required, then a precise knowledge of the monochromatic radiation
properties is necessary.
Figure 3.7 shows the radiative power of an actual real body at T = 2,000 K. The results of Planck’s
distribution law, equation (3.4), for a black and a grey body are given as a comparison.

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

38 3. Heat transfer

Figure 3.7: Radiative heat flux of a black,


grey and real body for T = 2,000 K

3.2.6 Radiation from a diffuse surface and direction-dependent radiation


Emissivity data for different materials are usually determined by measurements in the entire hemi-
sphere. Most of the real surfaces used in practice emit different amounts in different directions so
that only ideal surfaces as that of the black or grey body can be considered diffuse, i.e. indepen-
dent of the direction.
Measurement results, as shown in Figure 3.8, show that for example conductors have direction-
dependent emissivity, which increases negligibly with increasing viewing angle  , yet by reaching
90° decreases rapidly. On the other hand, insulators possess nearly constant and relatively large
emissivity over a wide range of viewing angles.

3.2.7 Radiation transfer


3.2.7.1 Radiation flux
We have discussed so far the radiation properties necessary to describe the heat flow from a body.
If, in addition to the total emitted heat flow from a body, we are also interested in the heat exchange
between the body and nearby objects, then it is necessary to know the exact radiation amount in
direction of the nearby body objects as well as the radiation absorbed by the body.
To gain better insight, we imagine a hemisphere of radius r over a small surface element dA, po-
sitioned at the center of the hemisphere, such that the entire heat flow from the surface element
passes through the hemisphere.

3.2.7.2 Radiation transfer between two bodies


If radiation transfer between two diffusely radiating surfaces of two bodies with different tem-
peratures occurs, then the hotter body emits more heat to the colder one, but not the other way
around so that .the net heat flow is from the hotter to the colder body. In an arbitrary orientation in
space, heat dQ l~2 flows from the surface area dA1 of body 1 to the surface area dA2 of body 2
(see Figure 3.9).

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

3. Heat transfer 39

Figure 3.8: Emissivity as a function of the emitting direction of selected materials, according to Eckert

The view factor F12 is that portion of the radiation that originates from the area A1 and which hits
the surface area A2:

(3.13)

this leads to:

A1 \12 = A2 \21 (3.14)

and finally to:

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

40 3. Heat transfer

Figure 3.9: Radiation between two surfaces

(3.15)

for cases where only black bodies are involved. Then the radiative power consists only of emitted
radiation, for which the Stefan-Boltzmann law equation is valid. As an example, Figure 3.10 shows
how the view factors can be determined for radiation transfer between perpendicular plates.

Radiation transfer between two grey surfaces


If, as a special case, radiation transfer between two black bodies occurs, then knowing the view
factor and introducing Stefan-Boltzmann’s law we can readily estimate the relationship of the net

Figure 3.10: View factor of


the radiation transfer between
perpendicular, rectangular
plates, according to Holman

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

3. Heat transfer 41

radiation transfer as a function of the temperature of the bodies. For grey bodies on the other hand,
the surface brightness includes not only the emitted but also portions of the reflected and transmit-
ted radiation since, unlike black bodies, the incident radiation is not fully absorbed.
Thus, for the radiation transfer between two non-transmissive solids

(3.16)

where e.g.

(3.17)

3
The solution methods for these equations are normally rather complex. Yet, for a number of simple,
often encountered geometries, complete relationships for the radiation transfer can be derived.

Radiation transfer between two infinitely long grey plates


According to Figure 3.11:

(3.18)

With l + _ = 1 and applying Kirchhoff’s law, _ = ¡ one gets

(3.19)

and accordingly:

(3.20)

Figure 3.11: Radiation between two grey


surfaces

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

42 3. Heat transfer

(3.21)

which can also be written as a function of the body temperature using the Stefan-Boltzmann law:

(3.22)

Radiation transfer between two self-enclosed grey bodies


In this case (see Figure 3.12):

(3.23)

or using Kirchhoff’s law and the reciprocal rule

In the case of enclosed bodies, the view factor \12 has the value 1 and thus

(3.24)

Figure 3.12: Radiation between two enclosed


bodies rate

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

3. Heat transfer 43

For the heat emitted from body 2 we get additionally parts of the reflected radiation, which origi-
nate from the surface brightness of the body itself – in other words “the body sees itself”.

(3.25)

Knowing that one gets

(3.26)
3

The equations are used to determine the surface brightness of both bodies, which after substitut-
ing and several transformations give a relationship for the radiation transfer

(3.27)

3.2.8 Gaseous radiation


Heat transfer between a gas volume and surfaces is considerably more complex compared to
heat transfer between two surfaces. In contrast to most solid bodies, gases are transparent. If
they absorb radiation, then the absorption is limited to a certain wavelength. The most relevant
components for gaseous radiation are the combustion products H2O and CO2 as well as some
hydrocarbons.

3.3 Heat conduction


This section deals with the simplest transport mechanism – heat conduction, excluding the discus-
sion on the additional combined influence of convection and radiation. The thermal conductivity
h [W/mK], already known from previous chapters, defines the material properties so that knowing
the molecular structure of the body or fluid transporting heat through conduction is not necessary.
It will suffice if we assume the medium to be homogeneous in structure.

3.3.1 Differential equations of the temperature field


We will examine the following experiment: if the surface temperature of a flat plate with thickness b
is changed by heating or cooling to the temperature T1 or T2, respectively,
. then a linear relationship
(see Figure 3.13) between the rate of heat flow per unit area Q /A and the temperature difference
(T1-T2) per plate thickness b can be observed.
The experiment above thus shows that a general relationship, determined by Fourier in 1822, can
be also expressed in the form

(Fourier’s equation) (3.28)

where ,T/,n is the temperature gradient perpendicular to a surface of constant temperature.

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

44 3. Heat transfer

Figure 3.13: Heat flow rate per unit area as a function of the temperature difference

The negative sign, as implied by the second law of thermodynamics, indicates that heat can only
flow from regions of higher to regions of lower temperature.

3.3.2 Steady state, one-dimensional heat conduction


3.3.2.1 Plane walls with given surface temperatures
For steady state, one-dimensional problems, without heat sources and constant heat conductivity h,
the Fourier equation simplifies to

The heat flux can be determined:

(3.29)

In case of a composite wall, i.e. consisting of many layers with different thickness and materials,
the equation must be calculated for each section in turn. The heat entering section 1 flows un-
changed out of section 3 for steady state, one-dimensional cases (see Figure 3.14).

Figure 3.14: Conduction through a composite wall

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

3. Heat transfer 45

It follows:

(3.30)

3.3.2.2 Plane walls with convective heat transfer


Although this chapter discusses only conduction heat transfer, we deal in this section with heat
transfer through convection since the latter is of importance and related by its boundary conditions 3
to conduction.
It has already been mentioned in the introduction that the energy transport from a flowing fluid to
the neighboring wall is not solely through conduction.
Only in the direct vicinity of the wall, where due to the viscosity of the fluid the velocity can be
neglected, we can calculate the heat as usual

(3.31)

using the temperature gradient in the fluid and the thermal conductivity hfl as parameters.

3.3.3 Unsteady state heat conduction without heat sources


Previous chapters have dealt with conduction processes for which a balance state has been
reached. The obtained temperature field was thus not a function of time. Since a general solution
of these equations is not possible, we can only describe typical examples used in practice.

3.3.3.1 Bodies with high values of thermal conductivity


If a body has a high thermal conductivity so that the thermal resistance of the body is small com-
pared to the heat transfer resistance between the body and the surrounding fluid, then a nearly
homogenous temperature will be established at each point in time during heating or cooling.
Hence, it is not necessary to implement the differential equation for the temperature field, equa-
tion (3.7). It is much easier to formulate the energy balance for the entire body, which states that
the inner energy of the body, which is a function of the time, will change due to heat transfer by
convection from the surface to the surroundings (see Figure 3.15).
If we define the dimensionless temperature e* = (T-T0) / (Tu – T0), where T is the temperature as a
function of the time and T0 is the initial temperature of the body, we get:

and thus

(3.32)

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

46 3. Heat transfer

Figure 3.15: Heating of a body

This differential equation can be integrated using the boundary condition (t = 0, T = T0, i.e. e* = 0)
for 0 ) e* ) 1 which finally gives

(3.33)

We can express the equation (3.33) introducing two dimensionless numbers, namely the Biot
number

Bi = _ L / h (3.34)

and the Fourier number

Fo = h t/l c L2 = a t/L2 (3.35)

where L is the characteristic length, resulting from the ratio V/A, in the form of:

(3.36)

In other words, the temperature history of a body (Figure 3.16) with high thermal conductivity
and its surrounding is in general the same for all problems where the boundary conditions are
described by equal Biot numbers.
Moreover, this is valid for all problems where the Biot number is very small, Bi << 1. Equation (3.36)
may still be used as a first approximation of the heat transfer behavior of a body in cases where
the above conditions do not apply.
In cases where the thermal resistance within a body cannot be neglected, a general, analytical
solution is possible only after complex calculations and only for a limited number of geometries
and boundary conditions.

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

3. Heat transfer 47

Figure 3.16: The temperature of


a body as a function of the time

3
3.4 Convection
We discussed the energy transport by heat conduction in chapter 3.3. A key role is assigned to the
molecular properties, described by the thermal conductivity of the material.
In some of the energy transport examples discussed previously we had to formulate the bound-
ary conditions for heat transfer from the body to the flowing fluid along the body. Without detailed
analysis of the physical relationships, we assumed that the heat transfer coefficient _ was as-
sumed to be known. In this section we will discuss the basic principles of this transport mechanism
(see Figure 3.17).
The major difference between the transport mechanisms of heat conduction and heat convection
can be demonstrated by a simple experiment, where a hot, horizontal wall is allowed to cool down
(Figure 3.18).
In the first experiment, the wall is cooled from below by a cold fluid, liquid or gas. For this arrange-
ment we have a stagnant fluid and the heat is transported solely by conduction to the area of lower
heat.

Figure 3.17: Heat transfer at a surface

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

48 3. Heat transfer

Figure 3.18: Cooling of a hot surface by conduction and convection

If we change the arrangement in such a way that the hot wall is placed at the bottom, then the fluid
will become unstable and the warmer areas (in the lower part) will rise up due to their lower density
and hence flow in the fluid will be initiated. This so-called natural convection amplifies the heat
transfer. In the third case shown (right most diagram), the hot wall is cooled by a medium, which
is pumped through the system. As was the case of natural convection, once again, in this case of
forced convection, it is not only the thermal conductivity of the fluid that determines the charac-
teristics of the heat transfer. An important, additional parameter that dictates the heat transfer is
obviously the velocity of the fluid, which depends both on the density differences as well as the
electrical power of the pump.

3.4.1 Application of the dimensional analysis for heat transfer


With the significant developments in computer technology, major advances have been made in the
numerical calculation of heat transfer. Nonetheless, it is still necessary for the major part of techni-
cal cases to use empirical heat transfer laws for design purposes.
These laws are a result of measurements and are shown graphically or in a form of empirical
formulas.
As will be shown in this section, it is not necessary to repeat the experiments if the fluid, geometry
or the flow conditions are changed.
The analytical solutions for heat transfer for forced and natural convection, dealt with previously,
have shown that the results can be represented in an appropriate form by a few dimensionless
numbers
Nu = Nu (Re, Pr) for forced convection
Nu = Nu (Gr, Pr) for natural convection
The solutions render not only the dimensionless numbers that describe the process but also their
functional relationship.
The heat transfer from a wall with the surface area A and temperature Tw to a fluid with the tem-
perature Tfl is:

Q/A = _ (Tfl – Tw) (3.37)

To determine the heat transfer coefficient _, a Nusselt number correlation is given in the next
chapter exemplary.

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

3. Heat transfer 49

The dimensionless numbers are defined as:

Nusselt number (3.38)

Reynolds number (3.39)

Grashof number (3.40)


3

Prandtl number (3.41)

The applicability of the given correlations should


be checked for each case in accordance with the
following criteria:
The geometry of the given problem must be analo-
gous to the geometry of the experiment from
which the heat transfer law is estimated.
Critical Reynolds or Grashof numbers are used
to determine whether the flow is laminar or tur-
bulent. A law valid for laminar flow cannot be ap-
plied for turbulent flow.

3.4.2 Cylinders in a flow parallel and per-


pendicular to their longitudinal axis
As long as the diameter of the body is much
greater than the thickness of the boundary layer,
cylinders in longitudinal flow can be regarded as
flat plates.
The local Nusselt number depicted in Figure 3.19
shows a strong influence of the circumference
angle which is a result of the transition from lami-
nar to turbulent flow and of the detachment of
the flow.

Figure 3.19: Local Nusselt number for cylinders


in a flow perpendicular to their longitudinal axis,
according to Holman

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015
to Table of Content

50 3. Heat transfer

Mean heat transfer


Nusselt law for the mean heat transfer

(3.42)

with the constant C relevant for the given ranges of Reynolds numbers and m (Table 3.3).

Table 3.3: Constants C and m for Reynolds number ranges

Red C m
0.4 – 4 0.989 0.330
4 – 40 0.911 0.385
40 – 4,000 0.683 0.466
4,000 – 40,000 0.193 0.618
40,000 – 400,000 0.0266 0.805

The functional relationship of equation (3.42) is shown in Figure 3.20.


For all Reynolds numbers the following simple approximation is valid (Whitaker, 1976)

(3.43)

Figure 3.20: Mean Nusselt number for circular cylinders in a flow perpendicular to their longitudinal axis

Einzellizenz für: Ulrich Renz - ulrich.renz@rwth-aachen.de


© DIV Deutscher Industrieverlag GmbH / Vulkan-Verlag GmbH - 17.11.2015

View publication stats

You might also like