Sie sind auf Seite 1von 34

CHAPTER 1:Thermal conductivity

1.1 Definition
1.1.2 Simple definition
1.1.3 General definition
1.2 Other quantities
1.3 Units
1.4 Measurement
1.5 Influencing factors
1.5.1 Temperature

1.5.2 Chemical phase

1.5.3 Thermal anisotropy

1.5.4 Electrical conductivity

1.5.5 Gaseous phases

1.5.6 Isotopic purity

1.6 Theoretical prediction


1.6.1 Gases

1.6.2 Liquids

1.6.3 Metals

1.6.4 Lattice waves

1.7 Conversion from specific to absolute units, and vice versa


CHAPTER 2:Thermocouple
2.1 Principle of operation
2.2 Circuit construction
2.3 Metallurgical grades
2.4 Aging of thermocouples
2.5 Applications
2.5.1 Steel industry

2.5.2 Gas appliance safety

2.6 Manufacturing
2.7 Power production
2.8 Process plants
2.9 Thermocouple as vacuum gauge

CHAPTER 3:Thermal contact conductance


3.1 Definition
3.2 Factors influencing contact conductance
3.2.1 Contact pressure

3.2.2 Interstitial materials

3.2.3 Surface roughness, waviness and flatness

3.2.4 Surface deformations

3.2.5 Surface cleanliness

3.3 Measurement of thermal contact conductance


CHAPTER 4:Copper in heat exchangers
4.1 INTRODUCTION
4.2 History
4.3 Beneficial properties of copper heat exchangers
4.3.1 Thermal conductivity

4.3.2 Corrosion resistance

4.3.3 Antimicrobial properties

4.3.4 Ease of inner grooving

4.4 Common applications for copper heat exchangers


4.4.1 Industrial facilities and power plants

4.4.2 Solar thermal water systems

4.4.3 HVAC systems

4.4.4 Gas water heaters

4.4.5 Forced air heating and cooling

4.4.6 Direct Exchange (DX) Geothermal Heating/Cooling

4.4.7 Electronic systems

4.5 New technologies


4.5.1 CuproBraze

4.5.2 Internally Grooved


CHAPTER 5: Thermal Conductivity Of Metal Rod
Aim
Theory
Description Of The Apparatus
Specifications
Prodedure
Observation Table
Calculation
Result
Precautions
CHAPTER 1:
Thermal conductivity
The thermal conductivity of a material is a measure of its ability to conduct
heat. It is commonly denoted by k, ʎ, or K,

Heat transfer occurs at a lower rate in materials of low thermal conductivity than
in materials of high thermal conductivity. For instance, metals typically have high
thermal conductivity and are very efficient at conducting heat, while the opposite
is true for insulating materials like Styrofoam. Correspondingly, materials of high
thermal conductivity are widely used in heat sink applications, and materials of
low thermal conductivity are used as thermal insulation. The reciprocal of
thermal conductivity is called thermal resistivity.

1.1 Definition
1.1.2 Simple definition
Consider a solid material placed between two environments of different
temperatures. Let be the temperature at X =0 and T2 the temperature at X=2,
and suppose T2 > T1. A possible realization of this scenario is a building on a cold
winter day: the solid material in this case would be the building wall, separating
the cold outdoor environment from the warm indoor environment.

According to the second law of thermodynamics, heat will flow from the hot
environment to the cold one in an attempt to equalize the temperature
difference. This is quantified in terms of a heat flux q, which gives the rate, per
unit area, at which heat flows in a given direction (in this case the x-direction). In
many materials, q is observed to be directly proportional to the temperature
difference and inversely proportional to the separation
−𝐤(𝐓𝟐−𝐓𝟏)
𝒒=
𝐋

The constant of proportionality k is the thermal conductivity; it is a physical


property of the material. In the present scenario, since T2 > T1 heat flows in the
minus x-direction and q is negative, which in turn means that k = 0. In general, k
is always defined to be positive. The same definition of k can also be extended to
gases and liquids, provided other modes of energy transport, such as convection
and radiation, are eliminated.

1.1.3 General definition


Thermal conduction is defined as the transport of energy due to random
molecular motion across a temperature gradient. It is distinguished from energy
transport by convection and molecular work in that it does not involve
macroscopic flows or work-performing internal stresses.

Energy flow due to thermal conduction is classified as heat and is quantified by


the vector q(r,t), which gives the heat flux at position r and time t . According to
the second law of thermodynamics, heat flows from high to low temperature.
Hence, it is reasonable to postulate that q(r,t), is proportional to the gradient of
the temperature field T(r,t)

1.2 Other quantities


In engineering practice, it is common to work in terms of quantities which are
derivative to thermal conductivity and implicitly take into account design-
specific features such as component dimensions. For instance, thermal
conductance is defined as the quantity of heat that passes in unit time through a
plate of particular area and thickness when its opposite faces differ in
temperature by one kelvin. For a plate of thermal conductivity k, area A, and
thickness L, the conductance is Ka/L, measured in W⋅ K−1. The relationship
between thermal conductivity and conductance is analogous to the relationship
between electrical conductivity and electrical conductance.

Thermal resistance is the inverse of thermal conductance. It is a convenient


measure to use in multicomponent design since thermal resistances are additive
when occurring in series.

There is also a measure known as the heat transfer coefficient: the quantity of
heat that passes per unit time through a unit area of a plate of particular
thickness when its opposite faces differ in temperature by one kelvin.In ASTM
C168-15, this area-independent quantity is referred to as the "thermal
conductance".The reciprocal of the heat transfer coefficient is thermal
insulance. In summary, for a plate of thermal conductivity k, area A, and
thickness L, we have

 thermal conductance = kA/L, measured in W·K−1.


o thermal resistance = L/(kA), measured in K·W−1.
 heat transfer coefficient = k/L, measured in W·K−1·m−2.
o thermal insulance = L/k, measured in K·m2·W−1.

The heat transfer coefficient is also known as thermal admittance in the sense
that the material may be seen as admitting heat to flow.

An additional term, thermal transmittance, quantifies the thermal conductance of


a structure along with heat transfer due to convection and radiation. It is
measured in the same units as thermal conductance and is sometimes known as
the composite thermal conductance. The term U-value is also used.

Finally, thermal diffusivity α, combines thermal conductivity with density and


specific heat

k
∝=
PCp

As such, it quantifies the thermal inertia of a material, i.e. the relative difficulty in
heating a material to a given temperature using heat sources applied at the
boundary.

1.3 Units
In the International System of Units (SI), thermal conductivity is measured in
watts per meter-kelvin (W/(m⋅K)). Some papers report in watts per centimeter-
kelvin (W/(cm⋅K)).

In imperial units, thermal conductivity is measured in BTU/(h⋅ft⋅°F).

The dimension of thermal conductivity is M1L1T−3Θ−1, expressed in terms of


the dimensions mass (M), length (L), time (T), and temperature (Θ).

Other units which are closely related to the thermal conductivity are in common
use in the construction and textile industries. The construction industry makes
use of units such as the R-value (resistance) and the U-value (transmittance).
Although related to the thermal conductivity of a material used in an insulation
product, R- and U-values are dependent on the thickness of the product. Likewise
the textile industry has several units including the tog and the clo which express
thermal resistance of a material in a way analogous to the R-values used in the
construction industry.
1.4 Measurement
There are several ways to measure thermal conductivity; each is suitable for a
limited range of materials. Broadly speaking, there are two categories of
measurement techniques: steady-state and transient. Steady-state techniques
infer the thermal conductivity from measurements on the state of a material once
a steady-state temperature profile has been reached, whereas transient
techniques operate on the instantaneous state of a system during the approach to
steady state. Lacking an explicit time component, steady-state techniques do not
require complicated signal analysis (steady state implies constant signals). The
disadvantage is that a well-engineered experimental setup is usually needed, and
the time required to reach steady state precludes rapid measurement.

In comparison with solid materials, the thermal properties of fluids are more
difficult to study experimentally. This is because in addition to thermal
conduction, convective and radiative energy transport are usually present unless
measures are taken to limit these processes. The formation of an insulating
boundary layer can also result in an apparent reduction in the thermal
conductivity.

1.5 Influencing factors


1.5.1 Temperature
The effect of temperature on thermal conductivity is different for metals and
nonmetals. In metals, heat conductivity is primarily due to free electrons.
Following the Wiedemann–Franz law, thermal conductivity of metals is
approximately proportional to the absolute temperature (in kelvins) times
electrical conductivity. In pure metals the electrical conductivity decreases with
increasing temperature and thus the product of the two, the thermal
conductivity, stays approximately constant. However, as temperatures approach
absolute zero, the thermal conductivity decreases sharply. In alloys the change in
electrical conductivity is usually smaller and thus thermal conductivity increases
with temperature, often proportionally to temperature. Many pure metals have a
peak thermal conductivity between 2 K and 10 K.

On the other hand, heat conductivity in nonmetals is mainly due to lattice


vibrations (phonons). Except for high-quality crystals at low temperatures, the
phonon mean free path is not reduced significantly at higher temperatures. Thus,
the thermal conductivity of nonmetals is approximately constant at high
temperatures. At low temperatures well below the Debye temperature, thermal
conductivity decreases, as does the heat capacity, due to carrier scattering from
defects at very low temperatures.

1.5.2 Chemical phase


When a material undergoes a phase change (e.g. from solid to liquid), the thermal
conductivity may change abruptly. For instance, when ice melts to form liquid
water at 0 °C, the thermal conductivity changes from 2.18 W/(m⋅K) to
0.56 W/(m⋅K). Even more dramatically, the thermal conductivity of a fluid
diverges in the vicinity of the vapor-liquid critical point.

1.5.3 Thermal anisotropy


Some substances, such as non-cubic crystals, can exhibit different thermal
conductivities along different crystal axes, due to differences in phonon coupling
along a given crystal axis. Sapphire is a notable example of variable thermal
conductivity based on orientation and temperature, with 35 W/(m⋅K) along the c
axis and 32 W/(m⋅K) along the a axis.[23] Wood generally conducts better along
the grain than across it. Other examples of materials where the thermal
conductivity varies with direction are metals that have undergone heavy cold
pressing, laminated materials, cables, the materials used for the Space Shuttle
thermal protection system, and fiber-reinforced composite structures. When
anisotropy is present, the direction of heat flow may not be exactly the same as
the direction of the thermal gradient.

1.5.4 Electrical conductivity


In metals, thermal conductivity approximately tracks electrical conductivity
according to the Wiedemann–Franz law, as freely moving valence electrons
transfer not only electric current but also heat energy. However, the general
correlation between electrical and thermal conductance does not hold for other
materials, due to the increased importance of phonon carriers for heat in non-
metals. Highly electrically conductive silver is less thermally conductive than
diamond, which is an electrical insulator, but due to its orderly array of atoms it
is conductive of heat via phonons.

1.5.5 Gaseous phases

Exhaust system components with ceramic coatings having a low thermal


conductivity reduce heating of nearby sensitive components

Air and other gases are generally good insulators, in the absence of convection.
Therefore, many insulating materials function simply by having a large number
of gas-filled pockets which obstruct heat conduction pathways. Examples of these
include expanded and extruded polystyrene (popularly referred to as
"styrofoam") and silica aerogel, as well as warm clothes. Natural, biological
insulators such as fur and feathers achieve similar effects by trapping air in
pores, pockets or voids, thus dramatically inhibiting convection of air or water
near an animal's skin.

Low density gases, such as hydrogen and helium typically have high thermal
conductivity. Dense gases such as xenon and dichlorodifluoromethane have low
thermal conductivity. An exception, sulfur hexafluoride, a dense gas, has a
relatively high thermal conductivity due to its high heat capacity. Argon and
krypton, gases denser than air, are often used in insulated glazing (double paned
windows) to improve their insulation characteristics.

The thermal conductivity through bulk materials in porous or granular form is


governed by the type of gas in the gaseous phase, and its pressure. At lower
pressures, the thermal conductivity of a gaseous phase is reduced, with this
behaviour governed by the Knudsen number, defined as Kn=l/d, where l is the
mean free path of gas molecules and d is the typical gap size of the space filled by
the gas. In a granular material d corresponds to the characteristic size of the
gaseous phase in the pores or intergranular spaces.

1.5.6 Isotopic purity


The thermal conductivity of a crystal can depend strongly on isotopic purity,
assuming other lattice defects are negligible. A notable example is diamond: at a
temperature of around 100 K the thermal conductivity increases from 10,000
W·m−1·K−1 for natural type IIa diamond (98.9% 12C), to 41,000 for 99.9%
enriched synthetic diamond. A value of 200,000 is predicted for 99.999% 12C at
80 K, assuming an otherwise pure crystal.

1.6 Theoretical prediction


The atomic mechanisms of thermal conduction vary among different materials,
and in general depend on details of the microscopic structure and atomic
interactions. As such, thermal conductivity is difficult to predict from first-
principles. Any expressions for thermal conductivity which are exact and general,
e.g. the Green-Kubo relations, are difficult to apply in practice, typically
consisting of averages over multiparticle correlation functions. A notable
exception is a dilute gas, for which a well-developed theory exists expressing
thermal conductivity accurately and explicitly in terms of molecular parameters.

In a gas, thermal conduction is mediated by discrete molecular collisions. In a


simplified picture of a solid, thermal conduction occurs by two mechanisms: 1)
the migration of free electrons and 2) lattice vibrations (phonons). The first
mechanism dominates in pure metals and the second in non-metallic solids. In
liquids, by contrast, the precise microscopic mechanisms of thermal conduction
are poorly understood.

1.6.1 Gases
In a simplified model of a dilute monatomic gas, molecules are modeled as rigid
spheres which are in constant motion, colliding elastically with each other and
with the walls of their container. Consider such a gas at temperature and T with
density p, specific heat Cp and molecular mass m. Under these assumptions, an
elementary calculation yields for the thermal conductivity.

1.6.2 Liquids
The exact mechanisms of thermal conduction are poorly understood in liquids:
there is no molecular picture which is both simple and accurate. An example of a
simple but very rough theory is that of Bridgman, in which a liquid is ascribed a
local molecular structure similar to that of a solid, i.e. with molecules located
approximately on a lattice. Elementary calculations then lead to the expression.

1.6.3 Metals
For metals at low temperatures the heat is carried mainly by the free electrons.
In this case the mean velocity is the Fermi velocity which is temperature
independent. The mean free path is determined by the impurities and the crystal
imperfections which are temperature independent as well. So the only
temperature-dependent quantity is the heat capacity c, which, in this case, is
proportional to T.

1.6.4 Lattice waves


Heat transport in both amorphous and crystalline dielectric solids is by way of
elastic vibrations of the lattice (i.e., phonons). This transport mechanism is
theorized to be limited by the elastic scattering of acoustic phonons at lattice
defects. This has been confirmed by the experiments of Chang and Jones on
commercial glasses and glass ceramics, where the mean free paths were found to
be limited by "internal boundary scattering" to length scales of 10−2 cm to
10−3 cm.

1.7 Conversion from specific to absolute units, and vice


versa
Specific thermal conductivity is a materials property used to compare the heat-
transfer ability of different materials (i.e., an intensive property). Absolute
thermal conductivity, in contrast, is a component property used to compare the
heat-transfer ability of different components (i.e., an extensive property).
Components, as opposed to materials, take into account size and shape, including
basic properties such as thickness and area, instead of just material type. In this
way, thermal-transfer ability of components of the same physical dimensions, but
made of different materials, may be compared and contrasted, or components of
the same material, but with different physical dimensions, may be compared and
contrasted. In component datasheets and tables, since actual, physical
components with distinct physical dimensions and characteristics are under
consideration, thermal resistance is frequently given in absolute units of K/W or
°C/W, since the two are equivalent. However, thermal conductivity, which is its
reciprocal, is frequently given in specific units of W/(K.m). It is therefore often
necessary to convert between absolute and specific units, by also taking a
component's physical dimensions into consideration, in order to correlate the
two using information provided, or to convert tabulated values of specific
thermal conductivity into absolute thermal resistance values for use in thermal
resistance calculations. This is particularly useful, for example, when calculating
the maximum power a component can dissipate as heat, as demonstrated in the
example calculation.
CHAPTER 2:
Thermocouple
A thermocouple is an electrical device consisting of two dissimilar electrical
conductors forming an electrical junction. A thermocouple produces a
temperature-dependent voltage as a result of the thermoelectric effect, and this
voltage can be interpreted to measure temperature. Thermocouples are a widely
used type of temperature sensor.

Commercial thermocouples are inexpensive,[2] interchangeable, are supplied


with standard connectors, and can measure a wide range of temperatures. In
contrast to most other methods of temperature measurement, thermocouples
are self powered and require no external form of excitation. The main limitation
with thermocouples is precision; system errors of less than one degree Celsius
(°C) can be difficult to achieve.

Thermocouples are widely used in science and industry. Applications include


temperature measurement for kilns, gas turbine exhaust, diesel engines, and
other industrial processes. Thermocouples are also used in homes, offices and
businesses as the temperature sensors in thermostats, and also as flame sensors
in safety devices for gas-powered appliances.

2.1 Principle of operation

In 1821, the German physicist Thomas Johann Seebeck discovered that when
different metals are joined at the ends and there is a temperature difference
between the joints, a magnetic field is observed. At the time, Seebeck referred to
this consequence as thermo-magnetism. The magnetic field he observed was
later shown to be due to thermo-electric current. In practical use, the voltage
generated at a single junction of two different types of wire is what is of interest
as this can be used to measure temperature at very high and low temperatures.
The magnitude of the voltage depends on the types of wire being used. Generally,
the voltage is in the microvolt range and care must be taken to obtain a usable
measurement. Although very little current flows, power can be generated by a
single thermocouple junction. Power generation using multiple thermocouples,
as in a thermopile, is common.
The standard configuration for thermocouple usage is shown in the figure.
Briefly, the desired temperature Tsense is obtained using three inputs—the
characteristic function E(T) of the thermocouple, the measured voltage V, and the
reference junctions' temperature Tref. The solution to the equation E(Tsense) = V +
E(Tref) yields Tsense. These details are often hidden from the user since the
reference junction block (with Tref thermometer), voltmeter, and equation solver
are combined into a single product.

2.2 Circuit construction


A common error in thermocouple construction is related to cold junction
compensation. If an error is made on the estimation of Tref, the same error will be
carried over to the temperature measurement. For the simplest measurements,
thermocouple wires are connected to copper far away from the hot or cold point
whose temperature is measured; the cold junction is then assumed to be at room
temperature, but that temperature can vary.

Junctions should be made in a reliable manner, but there are many possible
approaches to accomplish this. For low temperatures, junctions can be brazed or
soldered; however, it may be difficult to find a suitable flux and this may not be
suitable at the sensing junction due to the solder's low melting point. Reference
and extension junctions are therefore usually made with screw terminal blocks.
For high temperatures, a common approach is a spot weld or crimp using a
durable material.

A common myth regarding thermocouples is that junctions must be made cleanly


without involving a third metal, to avoid unwanted added EMFs. This may result
from another common misunderstanding that the voltage is generated at the
junction. In fact, the junctions should in principle have uniform internal
temperature; therefore, no voltage is generated at the junction. The voltage is
generated in the thermal gradient, along the wire.

A thermocouple produces small signals, often microvolts in magnitude. Precise


measurements of this signal require an amplifier with low input offset voltage
and with care taken to avoid thermal EMFs from self-heating within the
voltmeter itself. If the thermocouple wire has a high resistance for some reason
(poor contact at junctions, or very thin wires used for fast thermal response), the
measuring instrument should have high input impedance to prevent an offset in
the measured voltage. A useful feature in thermocouple instrumentation will
simultaneously measure resistance and detect faulty connections in the wiring or
at thermocouple junctions.

2.3 Metallurgical grades


While a thermocouple wire type is often described by its chemical composition,
the actual aim is to produce a pair of wires that follow a standardized E(T)curve.
Impurities affect each batch of metal differently, producing variable Seebeck
coefficients. To match the standard behaviour, thermocouple wire manufacturers
will deliberately mix in additional impurities to "dope" the alloy, compensating
for uncontrolled variations in source material. As a result, there are standard and
specialized grades of thermocouple wire, depending on the level of precision
demanded in the thermocouple behaviour. Precision grades may only be
available in matched pairs, where one wire is modified to compensate for
deficiencies in the other wire.

A special case of thermocouple wire is known as "extension grade", designed to


carry the thermoelectric circuit over a longer distance. Extension wires follow
the stated E(T) curve but for various reasons they are not designed to be used in
extreme environments and so they cannot be used at the sensing junction in
some applications. For example, an extension wire may be in a different form,
such as highly flexible with stranded construction and plastic insulation, or be
part of a multi-wire cable for carrying many thermocouple circuits. With
expensive noble metal thermocouples, the extension wires may even be made of
a completely different, cheaper material that mimics the standard type over a
reduced temperature range.

2.4 Aging of thermocouples


Thermocouples are often used at high temperatures and in reactive furnace
atmospheres. In this case, the practical lifetime is limited by thermocouple aging.
The thermoelectric coefficients of the wires in a thermocouple that is used to
measure very high temperatures may change with time, and the measurement
voltage accordingly drops. The simple relationship between the temperature
difference of the junctions and the measurement voltage is only correct if each
wire is homogeneous (uniform in composition). As thermocouples age in a
process, their conductors can lose homogeneity due to chemical and
metallurgical changes caused by extreme or prolonged exposure to high
temperatures. If the aged section of the thermocouple circuit is exposed to a
temperature gradient, the measured voltage will differ, resulting in error.

Aged thermocouples are only partly modified; for example, being unaffected in
the parts outside the furnace. For this reason, aged thermocouples cannot be
taken out of their installed location and recalibrated in a bath or test furnace to
determine error. This also explains why error can sometimes be observed when
an aged thermocouple is pulled partly out of a furnace—as the sensor is pulled
back, aged sections may see exposure to increased temperature gradients from
hot to cold as the aged section now passes through the cooler refractory area,
contributing significant error to the measurement. Likewise, an aged
thermocouple that is pushed deeper into the furnace might sometimes provide a
more accurate reading if being pushed further into the furnace causes the
temperature gradient to occur only in a fresh section.
2.5 Applications
Thermocouples are suitable for measuring over a large temperature range, from
−270 up to 3000 °C (for a short time, in inert atmosphere). Applications include
temperature measurement for kilns, gas turbine exhaust, diesel engines, other
industrial processes and fog machines. They are less suitable for applications
where smaller temperature differences need to be measured with high accuracy,
for example the range 0–100 °C with 0.1 °C accuracy. For such applications
thermistors, silicon bandgap temperature sensors and resistance thermometers
are more suitable.

2.5.1 Steel industry


Type B, S, R and K thermocouples are used extensively in the steel and iron
industries to monitor temperatures and chemistry throughout the steel making
process. Disposable, immersible, type S thermocouples are regularly used in the
electric arc furnace process to accurately measure the temperature of steel
before tapping. The cooling curve of a small steel sample can be analyzed and
used to estimate the carbon content of molten steel.

2.5.2 Gas appliance safety


Many gas-fed heating appliances such as ovens and water heaters make use of a
pilot flame to ignite the main gas burner when required. If the pilot flame goes
out, unburned gas may be released, which is an explosion risk and a health
hazard. To prevent this, some appliances use a thermocouple in a fail-safe circuit
to sense when the pilot light is burning. The tip of the thermocouple is placed in
the pilot flame, generating a voltage which operates the supply valve which feeds
gas to the pilot. So long as the pilot flame remains lit, the thermocouple remains
hot, and the pilot gas valve is held open. If the pilot light goes out, the
thermocouple temperature falls, causing the voltage across the thermocouple to
drop and the valve to close.

Where the probe may be easily placed above the flame, a rectifying sensor may
often be used instead. With part ceramic construction, they may also be known as
flame rods, flame sensors or flame detection electrodes.

Some combined main burner and pilot gas valves (mainly by Honeywell) reduce
the power demand to within the range of a single universal thermocouple heated
by a pilot (25 mV open circuit falling by half with the coil connected to a 10–12
mV, 0.2–0.25 A source, typically) by sizing the coil to be able to hold the valve
open against a light spring, but only after the initial turning-on force is provided
by the user pressing and holding a knob to compress the spring during lighting of
the pilot. These systems are identifiable by the "press and hold for x minutes" in
the pilot lighting instructions. (The holding current requirement of such a valve
is much less than a bigger solenoid designed for pulling the valve in from a closed
position would require.) Special test sets are made to confirm the valve let-go
and holding currents, because an ordinary milliammeter cannot be used as it
introduces more resistance than the gas valve coil. Apart from testing the open
circuit voltage of the thermocouple, and the near short-circuit DC continuity
through the thermocouple gas valve coil, the easiest non-specialist test is
substitution of a known good gas valve.

Some systems, known as millivolt control systems, extend the thermocouple


concept to both open and close the main gas valve as well. Not only does the
voltage created by the pilot thermocouple activate the pilot gas valve, it is also
routed through a thermostat to power the main gas valve as well. Here, a larger
voltage is needed than in a pilot flame safety system described above, and a
thermopile is used rather than a single thermocouple. Such a system requires no
external source of electricity for its operation and thus can operate during a
power failure, provided that all the other related system components allow for
this. This excludes common forced air furnaces because external electrical power
is required to operate the blower motor, but this feature is especially useful for
un-powered convection heaters. A similar gas shut-off safety mechanism using a
thermocouple is sometimes employed to ensure that the main burner ignites
within a certain time period, shutting off the main burner gas supply valve should
that not happen.

Out of concern about energy wasted by the standing pilot flame, designers of
many newer appliances have switched to an electronically controlled pilot-less
ignition, also called intermittent ignition. With no standing pilot flame, there is no
risk of gas buildup should the flame go out, so these appliances do not need
thermocouple-based pilot safety switches. As these designs lose the benefit of
operation without a continuous source of electricity, standing pilots are still used
in some appliances. The exception is later model instantaneous (aka "tankless")
water heaters that use the flow of water to generate the current required to
ignite the gas burner; these designs also use a thermocouple as a safety cut-off
device in the event the gas fails to ignite, or if the flame is extinguished.

2.5.3 Thermopile radiation sensors


Thermopiles are used for measuring the intensity of incident radiation, typically
visible or infrared light, which heats the hot junctions, while the cold junctions
are on a heat sink. It is possible to measure radiative intensities of only a few
μW/cm2 with commercially available thermopile sensors. For example, some
laser power meters are based on such sensors; these are specifically known as
thermopile laser sensor.

The principle of operation of a thermopile sensor is distinct from that of a


bolometer, as the latter relies on a change in resistance.
2.6 Manufacturing
Thermocouples can generally be used in the testing of prototype electrical and
mechanical apparatus. For example, switchgear under test for its current
carrying capacity may have thermocouples installed and monitored during a heat
run test, to confirm that the temperature rise at rated current does not exceed
designed limits.

2.7 Power production


A thermocouple can produce current to drive some processes directly, without
the need for extra circuitry and power sources. For example, the power from a
thermocouple can activate a valve when a temperature difference arises. The
electrical energy generated by a thermocouple is converted from the heat which
must be supplied to the hot side to maintain the electric potential. A continuous
transfer of heat is necessary because the current flowing through the
thermocouple tends to cause the hot side to cool down and the cold side to heat
up (the Peltier effect).

Thermocouples can be connected in series to form a thermopile, where all the


hot junctions are exposed to a higher temperature and all the cold junctions to a
lower temperature. The output is the sum of the voltages across the individual
junctions, giving larger voltage and power output. In a radioisotope
thermoelectric generator, the radioactive decay of transuranic elements as a heat
source has been used to power spacecraft on missions too far from the Sun to use
solar power.

Thermopiles heated by kerosene lamps were used to run batteryless radio


receivers in isolated areas. There are commercially produced lanterns that use
the heat from a candle to run several light-emitting diodes, and
thermoelectrically-powered fans to improve air circulation and heat distribution
in wood stoves.

2.8 Process plants


Chemical production and petroleum refineries will usually employ computers for
logging and for limit testing the many temperatures associated with a process,
typically numbering in the hundreds. For such cases, a number of thermocouple
leads will be brought to a common reference block (a large block of copper)
containing the second thermocouple of each circuit. The temperature of the block
is in turn measured by a thermistor. Simple computations are used to determine
the temperature at each measured location.

2.9 Thermocouple as vacuum gauge


A thermocouple can be used as a vacuum gauge over the range of approximately
0.001 to 1 torr absolute pressure. In this pressure range, the mean free path of
the gas is comparable to the dimensions of the vacuum chamber, and the flow
regime is neither purely viscous nor purely molecular. In this configuration, the
thermocouple junction is attached to the centre of a short heating wire, which is
usually energised by a constant current of about 5 mA, and the heat is removed at
a rate related to the thermal conductivity of the gas.

The temperature detected at the thermocouple junction depends on the thermal


conductivity of the surrounding gas, which depends on the pressure of the gas.
The potential difference measured by a thermocouple is proportional to the
square of pressure over the low- to medium-vacuum range. At higher (viscous
flow) and lower (molecular flow) pressures, the thermal conductivity of air or
any other gas is essentially independent of pressure. The thermocouple was first
used as a vacuum gauge by Voege in 1906. The mathematical model for the
thermocouple as a vacuum gauge is quite complicated, as explained in detail by
Van Atta,
CHAPTER 3:
Thermal contact conductance
In physics, thermal contact conductance is the study of heat conduction
between solid bodies in thermal contact. The thermal contact conductance
coefficient, hc, is a property indicating the thermal conductivity, or ability to
conduct heat, between two bodies in contact. The inverse of this property is
termed thermal contact resistance.

3.1 Definition

When two solid bodies come in contact, such as A and B in Figure 1, heat flows
from the hotter body to the colder body. From experience, the temperature
profile along the two bodies varies, approximately, as shown in the figure. A
temperature drop is observed at the interface between the two surfaces in
contact. This phenomenon is said to be a result of a thermal contact resistance
existing between the contacting surfaces. Thermal contact resistance is defined
as the ratio between this temperature drop and the average heat flow across the
interface.

3.2 Factors influencing contact conductance


Thermal contact conductance is a complicated phenomenon, influenced by many
factors. Experience shows that the most important ones are as follows:
3.2.1 Contact pressure
For thermal transport between two contacting bodies, such as particles in a
granular medium, the contact pressure is the factor of most influence on overall
contact conductance. As contact pressure grows, true contact area increases and
contact conductance grows (contact resistance becomes smaller). Since the
contact pressure is the most important factor, most studies, correlations and
mathematical models for measurement of contact conductance are done as a
function of this factor. The thermal contact resistance of certain sandwich kinds
of materials that are manufactured by rolling under high temperatures may
sometimes be ignored because the decrease in thermal conductivity between
them is negligible.

3.2.2 Interstitial materials


No truly smooth surfaces really exist, and surface imperfections are visible under
a microscope. As a result, when two bodies are pressed together, contact is only
performed in a finite number of points, separated by relatively large gaps, as can
be shown in Fig. 2. Since the actual contact area is reduced, another resistance for
heat flow exists. The gases/fluids filling these gaps may largely influence the total
heat flow across the interface. The thermal conductivity of the interstitial
material and its pressure, examined through reference to the Knudsen number,
are the two properties governing its influence on contact conductance, and
thermal transport in heterogeneous materials in general. In the absence of
interstitial materials, as in a vacuum, the contact resistance will be much larger,
since flow through the intimate contact points is dominant.

3.2.3 Surface roughness, waviness and flatness


One can characterise a surface that has undergone certain finishing operations by
three main properties of: roughness, waviness,and fractal dimension. Among
these, roughness and fractality are of most importance, with roughness often
indicated in terms of a rms value and surface fractality denoted generally by Df.
The effect of surface structures on thermal conductivity at interfaces is analogous
to the concept of electrical contact resistance, also known as ECR, involving
contact patch restricted transport of phonons rather than electrons.

3.2.4 Surface deformations


When the two bodies come in contact, surface deformation may occur on both
bodies. This deformation may either be plastic or elastic, depending on the
material properties and the contact pressure. When a surface undergoes plastic
deformation, contact resistance is lowered, since the deformation causes the
actual contact area to increase.
3.2.5 Surface cleanliness
The presence of dust particles, acids, etc., can also influence the contact
conductance.

3.3 Measurement of thermal contact conductance


Going back to Formula 2, calculation of the thermal contact conductance may
prove difficult, even impossible, due to the difficulty in measuring the contact
area, A (A product of surface characteristics, as explained earlier). Because of
this, contact conductance/resistance is usually found experimentally, by using a
standard apparatus.[9]

The results of such experiments are usually published in Engineering literature,


on journals such as Journal of Heat Transfer, International Journal of Heat and
Mass Transfer, etc. Unfortunately, a centralized database of contact conductance
coefficients does not exist, a situation which sometimes causes companies to use
outdated, irrelevant data, or not taking contact conductance as a consideration at
all.

CoCoE (Contact Conductance Estimator), a project founded to solve this problem


and create a centralized database of contact conductance data and a computer
program that uses it, was started in 2006.

3.4 Thermal boundary conductance


While a finite thermal contact conductance is due to voids at the interface,
surface waviness, and surface roughness, etc., a finite conductance exists even at
near ideal interfaces as well. This conductance, known as thermal boundary
conductance, is due to the differences in electronic and vibrational properties
between the contacting materials. This conductance is generally much higher
than thermal contact conductance, but becomes important in nanoscale material
systems.
CHAPTER 4:
Copper in heat exchangers
4.1 INTRODUCTION
Heat exchangers are devices that transfer heat to achieve desired heating or
cooling. An important design aspect of heat exchanger technology is the selection
of appropriate materials to conduct and transfer heat fast and efficiently.

Copper has many desirable properties for thermally efficient and durable heat
exchangers. First and foremost, copper is an excellent conductor of heat. This
means that copper's high thermal conductivity allows heat to pass through it
quickly. Other desirable properties of copper in heat exchangers include its
corrosion resistance, biofouling resistance, maximum allowable stress and
internal pressure, creep rupture strength, fatigue strength, hardness, thermal
expansion, specific heat, antimicrobial properties, tensile strength, yield strength,
high melting point, alloyability, ease of fabrication, and ease of joining.

The combination of these properties enable copper to be specified for heat


exchangers in industrial facilities, HVAC systems, vehicular coolers and radiators,
and as heat sinks to cool computers, disk drives, televisions, computer monitors,
and other electronic equipment.[1] Copper is also incorporated into the bottoms
of high-quality cookware because the metal conducts heat quickly and
distributes it evenly.

Non-copper heat exchangers are also available. Some alternative materials


include aluminium, carbon steel, stainless steel, nickel alloys, and titanium.

This article focuses on beneficial properties and common applications of copper


in heat exchangers. New copper heat exchanger technologies for specific
applications are also introduced.

4.2 History
Heat exchangers using copper and its alloys have evolved along with heat
transfer technologies over the past several hundred years. Copper condenser
tubes were first used in 1769 for steam engines. Initially, the tubes were made of
unalloyed copper. By 1870, Muntz metal, a 60% Cu-40% Zn brass alloy, was used
for condensers in seawater cooling. Admiralty metal, a 70% Cu-30% Zn yellow
brass alloy with 1% tin added to improve corrosion resistance, was introduced in
1890 for seawater service. By the 1920s, a 70% Cu-30% Ni alloy was developed
for naval condensers. Soon afterwards, a 2% manganese and 2% iron copper
alloy was introduced for better erosion resistance. A 90% Cu-10% Ni alloy first
became available in the 1950s, initially for seawater piping. This alloy is now the
most widely used copper-nickel alloy in marine heat exchangers.

Today, steam, evaporator, and condenser coils are made from copper and copper
alloys. These heat exchangers are used in air conditioning and refrigeration
systems, industrial and central heating and cooling systems, radiators, hot water
tanks, and under-floor heating systems.

Copper-based heat exchangers can be manufactured with copper


tube/aluminium fin, cupro-nickel, or all-copper constructions. Various coatings
can be applied to enhance corrosion resistance of the tubes and fins.

4.3 Beneficial properties of copper heat exchangers


4.3.1 Thermal conductivity
Thermal conductivity (k, also denoted as λ or κ) is a measure of a material's
ability to conduct heat. Heat transfer across materials of high thermal
conductivity occurs at a higher rate than across materials of low thermal
conductivity. In the International System of Units (SI), thermal conductivity is
measured in watts per meter Kelvin (W/(m•K)). In the Imperial System of
Measurement (British Imperial, or Imperial units), thermal conductivity is
measured in Btu/(hr•ft⋅F).

Copper has a thermal conductivity of 231 Btu/(hr-ft-F). This is higher than all
other metals except silver, a precious metal. Copper has a 60% better thermal
conductivity rating than aluminium and a 3,000% better rating than stainless
steel.

4.3.2 Corrosion resistance

Corrosion resistance is essential in heat transfer applications where


fluids are involved, such as in hot water tanks, radiators, etc. The only
affordable material that has similar corrosion resistance to copper is
stainless steel. However, the thermal conductivity of stainless steel is
1/30th that of copper. Aluminium tubes are not suitable for potable or
untreated water applications because it corrodes at pH<7.0 and
releases hydrogen gas.

Protective films can be applied to the inner surface of copper alloy


tubes to increase corrosion resistance. For certain applications, the
film is composed of iron. In power plant condensers, duplex tubes
consisting of an inner titanium layer with outer copper-nickel alloys
are employed. This enables the use of copper’s beneficial mechanical
and chemical properties (e.g., stress corrosion cracking, ammonia
attack) along with titanium’s excellent corrosion resistance. A duplex
tube with inner aluminium brass or copper-nickel and outer stainless
or mild steel can be used for cooling in the oil refining and
petrochemical industries.

4.3.3 Antimicrobial properties


Due to copper’s strong antimicrobial properties, copper fins can inhibit bacterial,
fungal and viral growths that commonly build up in air conditioning systems.
Hence, the surfaces of copper-based heat exchangers are cleaner for longer
periods of time than heat exchangers made from other metals. This benefit offers
a greatly expanded heat exchanger service life and contributes to improved air
quality. Heat exchangers fabricated separately from antimicrobial copper and
aluminium in a full-scale HVAC system have been evaluated for their ability to
limit microbial growth under conditions of normal flow rates using single-pass
outside air. Commonly used aluminium components developed stable biofilms of
bacteria and fungi within four weeks of operation. During the same time period,
antimicrobial copper was able to limit bacterial loads associated with the copper
heat exchanger fins by 99.99% and fungal loads by 99.74%.

Copper fin air conditioners have been deployed on buses in Shanghai to rapidly
and completely kill bacteria, viruses and fungi that were previously thriving on
non-copper fins and permitted to circulate around the systems. The decision to
replace aluminium with copper followed antimicrobial tests by the Shanghai
Municipal Center for Disease Control and Prevention (SCDC) from 2010 to 2012.
The study found that microbial levels on copper fin surfaces were significantly
lower than on aluminium, thereby helping to protect the health of bus
passengers.

Further information about the benefits of antimicrobial copper in HVAC systems


is available.

4.3.4 Ease of inner grooving


Internally grooved copper tube of smaller diameters is more thermally efficient,
materially efficient, and easier to bend and flare and otherwise work with. It is
generally easier to make inner grooved tubes out of copper, a very soft metal.

4.4 Common applications for copper heat exchangers


4.4.1 Industrial facilities and power plants
Copper alloys are extensively used as heat exchanger tubing in fossil and nuclear
steam generating electric power plants, chemical and petrochemical plants,
marine services, and desalination plants.
The largest use of copper alloy heat exchanger tubing on a per unit basis is in
utility power plants. These plants contain surface condensers, heaters, and
coolers, all of which contain copper tubing. The main surface condenser that
accepts turbine-steam discharges uses the most amount of copper.

Copper nickel is the group of alloys that are commonly specified in heat
exchanger or condenser tubes in evaporators of desalination plants, process
industry plants, air cooling zones of thermal power plants, high-pressure feed
water heaters, and sea water piping in ships.[11] The composition of the alloys can
vary from 90% Cu–10% Ni to 70% Cu–30% Ni.

Condenser and heat exchanger tubing of arsenical admiralty brass (Cu-Zn-Sn-As)


once dominated the industrial facility market. Aluminium brass later rose in
popularity because of its enhanced corrosion resistance.[22] Today, aluminium-
brass, 90%Cu-10%Ni, and other copper alloys are widely used in tubular heat
exchangers and piping systems in seawater, brackish water and fresh water.
Aluminium-brass, 90% Cu-10% Ni and 70% Cu-30% Ni alloys show good
corrosion resistance in hot de-aerated seawater and in brines in multi-stage flash
desalination plants.

Fixed tube liquid-cooled heat exchangers especially suitable for marine and
harsh applications can be assembled with brass shells, copper tubes, brass
baffles, and forged brass integral end hubs.

Copper alloy tubes can be supplied either with a bright metallic surface (CuNiO)
or with a thin, firmly attached oxide layer (aluminium brass). These finish types
allow for the formation of a protective layer.[24] The protective oxide surface is
best achieved when the system is operated for several weeks with clean, oxygen
containing cooling water. While the protective layer forms, supportive measures
can be carried out to enhance the process, such as the addition of iron sulfate or
intermittent tube cleaning. The protective film that forms on Cu-Ni alloys in
aerated seawater becomes mature in about three months at 60 °F and becomes
increasingly protective with time. The film is resistant to polluted waters,
irregular velocities, and other harsh conditions. Further details are available.

The biofouling resistance of Cu-Ni alloys enables heat exchange units to operate
for several months between mechanical cleanings. Cleanings are nevertheless
needed to restore original heat transfer capabilities. Chlorine injection can
extend the mechanical cleaning intervals to a year or more without detrimental
effects on the Cu-Ni alloys.

Further information about copper alloy heat exchangers for industrial facilities is
available.
4.4.2 Solar thermal water systems
Solar water heaters can be a cost-effective way to generate hot water for homes
in many regions of the world. Copper heat exchangers are important in solar
thermal heating and cooling systems because of copper's high thermal
conductivity, resistance to atmospheric and water corrosion, sealing and joining
by soldering, and mechanical strength. Copper is used both in receivers and in
primary circuits (pipes and heat exchangers for water tanks) of solar thermal
water systems.

Various types of solar collectors for residential applications are available with
either direct circulation (i.e., heats water and brings it directly to the home for
use) or indirect circulation (i.e., pumps a heat transfer fluid through a heat
exchanger, which then heats water that flows into the home) systems.[32] In an
evacuated tube solar hot water heater with an indirect circulation system, the
evacuated tubes contain a glass outer tube and metal absorber tube attached to a
fin. Solar thermal energy is absorbed within the evacuated tubes and is converted
into usable concentrated heat. Evacuated glass tubes have a double layer. Inside
the glass tube is the copper heat pipe. It is a sealed hollow copper tube that
contains a small amount of thermal transfer fluid (water or glycol mixture) which
under low pressure boils at a very low temperature. The copper heat pipe
transfers thermal energy from within the solar tube into a copper header. As the
solution circulates through the copper header, the temperature rises.

Other components in solar thermal water systems that contain copper include
solar heat exchanger tanks and solar pumping stations, along with pumps and
controllers.

4.4.3 HVAC systems


Air conditioning and heating in buildings and motor vehicles are two of the
largest applications for heat exchangers. While copper tube is used in most air
conditioning and refrigeration systems, typical air conditioning units currently
use aluminium fins. These systems can harbor bacteria and mold and develop
odors and fouling that can make them function poorly.[38] Stringent new
requirements including demands for increased operating efficiencies and the
reduction or elimination of harmful emissions are enhancing copper's role in
modern HVAC systems.

Copper’s antimicrobial properties can enhance the performance of HVAC systems


and associated indoor air quality. After extensive testing, copper became a
registered material in the U.S. for protecting heating and air conditioning
equipment surfaces against bacteria, mold, and mildew. Furthermore, testing
funded by the U.S. Department of Defense is demonstrating that all-copper air
conditioners suppress the growth of bacteria, mold and mildew that cause odors
and reduce system energy efficiency. Units made with aluminium have not been
demonstrating this benefit.
Copper can cause a galvanic reaction in the presence of other alloys, leading to
corrosion.

4.4.4 Gas water heaters


Water heating is the second largest energy use in the home. Gas-water heat
exchangers that transfer heat from gaseous fuels to water between 3 and 300
kilowatts thermal (kWth) have widespread residential and commercial use in
water heating and heating boiler appliance applications.

Demand is increasing for energy-efficient compact water heating systems.


Tankless gas water heaters produce hot water when needed. Copper heat
exchangers are the preferred material in these units because of their high
thermal conductivity and ease of fabrication. To protect these units in acidic
environments, durable coatings or other surface treatments are available. Acid-
resistant coatings are capable of withstanding temperatures of 1000 °C.

4.4.5 Forced air heating and cooling


Air-source heat pumps have been used for residential and commercial heating
and cooling for many years. These units rely on air-to-air heat exchange through
evaporator units similar to those used for air conditioners. Finned water to air
heat exchangers are most commonly used for forced air heating and cooling
systems, such as with indoor and outdoor wood furnaces, boilers, and stoves.
They can also be suitable for liquid cooling applications. Copper is specified in
supply and return manifolds and in tube coils.

4.4.6 Direct Exchange (DX) Geothermal Heating/Cooling


Geothermal heat pump technology, variously known as "ground source," "earth-
coupled," or "direct exchange," relies on circulating a refrigerant through buried
copper tubing for heat exchange. These units, which are considerably more
efficient than their air-source counterparts, rely on the constancy of ground
temperatures below the frost zone for heat transfer. The most efficient ground
source heat pumps use ACR, Type L or special-size copper tubing buried into the
ground to transfer heat to or from the conditioned space. Flexible copper tube
(typically 1/4-inch to 5/8-inch) can be buried in deep vertical holes, horizontally
in a relatively shallow grid pattern, in a vertical fence-like arrangement in
medium-depth trenches, or as custom configurations. Further information is
available.

4.4.7 Electronic systems


Copper and aluminium are used as heat sinks and heat pipes in electronic cooling
applications. A heat sink is a passive component that cools semiconductor and
optoelectronic devices by dissipating heat into the surrounding air. Heat sinks
have temperatures higher than their surrounding environments so that heat can
be transferred into the air by convection, radiation, and conduction.

Aluminium is the most prominently used heat sink material because of its lower
cost. Copper heat sinks are a necessity when higher levels of thermal
conductivity are needed. An alternative to all-copper or all-aluminium heat sinks
is the joining of aluminium fins to a copper base.

Copper heat sinks are die-cast and bound together in plates. They spread heat
quickly from the heat source to copper or aluminium fins and into the
surrounding air.

Heat pipes are used to move heat away from central processing units (CPUs) and
graphics processing units (GPUs) and towards heat sinks, where thermal energy
is dissipated into the environment. Copper and aluminium heat pipes are used
extensively in modern computer systems where increased power requirements
and associated heat emissions result in greater demands on cooling systems.

A heat pipe typically consists of a sealed pipe or tube at both the hot and cold
ends. Heat pipes utilize evaporative cooling to transfer thermal energy from one
point to another by the evaporation and condensation of a working fluid or
coolant. They are fundamentally better at heat conduction over larger distances
than heat sinks because their effective thermal conductivity is several orders of
magnitude greater than that of the equivalent solid conductor.

When it is desirable to maintain junction temperatures below 125–150 °C,


copper/water heat pipes are typically used. Copper/methanol heat pipes are
used if the application requires heat pipe operations below 0 °C.

4.5 New technologies


4.5.1 CuproBraze
CuproBraze is a copper-alloy heat exchanger technology developed for
applications that need to withstand harsh conditions. The technology is
particularly amenable for higher temperature and pressure environments
required in cleaner diesel engines that are being mandated by global
environmental regulations.

Applications for CuproBraze include charge air coolers, radiators, oil coolers,
climate control systems, and heat transfer cores.[51][52] CuproBraze is particularly
suited for charge air coolers and radiators in capital intensive industries where
machinery must operate for long periods of time under harsh conditions without
premature failures. For these reasons, CuproBraze is particularly suited for the
off-road vehicle, truck, bus, industrial engine, generator, locomotive, and military
equipment markets. The technology is also amenable for light trucks, SUVs and
passenger cars.
CuproBraze is an alternative to soldered copper/brass plate fin, soldered copper
brass serpentine fin, and brazed aluminium serpentine fin.[51] The technology
enables brazed copper serpentine fins to be used in copper-brass heat exchanger
designs. These are less expensive to manufacture than soldered serpentine fin
designs. They are also stronger, lighter, more durable, and have tougher joints.

4.5.2 Internally Grooved


The benefits of smaller-diameter internally grooved copper tube for heat transfer
are well documented.

Smaller diameter coils have better rates of heat transfer than conventional sized
coils and they can withstand higher pressures required by the new generation of
environmentally friendlier refrigerants. Smaller diameter coils also have lower
material costs because they require less refrigerant, fin, and coil materials; and
they enable the design of smaller and lighter high-efficiency air conditioners and
refrigerators because the evaporators and condensers coils are smaller and
lighter. MicroGroove uses a grooved inner surface of the tube to increase the
surface to volume ratio and increase turbulence to mix the refrigerant and
homogenize temperatures across the tube.
CHAPTER 5:
THERMAL CONDUCTIVITY OF METAL ROD
AIM:
To determine the thermal conductivity of given metal rod.
THEORY:
From Fourier’s law of heat conduction
−kA dT
𝑄=
dX

where, Q = Rate of heat conducted, W


A = Area of heat transfer, m²
k = Thermal conductivity of the material, W/m-K
dT / dx= Temperature gradient

Thermal conductivity is a property of the material and may be defined as the


amount of heat conducted per unit time through unit area, when a temperature
difference of unit degree is maintained across unit thickness.

DESCRIPTION OF THE APPARATUS:


The apparatus consists of Brass rod, one end of which is heated by an electric
heating coil while the other end projects into the cooling water jacket. The rod is
insulated with glass wool to minimize the radiation and convection loss from the
surface of the rod and thus ensure nearly constant temperature gradient
throughout the length of the rod. The temperature of the rod is measured at five
different locations. The heater is provided with a dimmerstat for controlling the
heat input. Water is circulated through the jacket and its flow rate and
temperature rise can be measured.
SPECIFICATIONS :

Specimen material : brass rod


Size of the Specimen : φ20 mm, 450mm long
Cylindrical shell : 300mm long
Voltmeter : Digital type, 0-300volt, AC
Ammeter : Digital type, 0-20amp, AC
Dimmer for heating Coil : 50-230v, 2amps
Heater : Band type Nichrome heater, 250 W
Thermocouple used : 11 nos.
Temperature indicator : Digital type, 0-400 , Cr-Al

PRODEDURE:
1. Power supply is given to the apparatus.
2. Give heat input to the heater by slowly rotating the dimmer and adjust the
voltage to say 50 V, 100 V, etc
3. Start the cooling water supply through the jacket and adjust its flow rate so
that the heat is taken away from the specimen constantly.
4. Allow sufficient time for the apparatus to reach steady state.
5. Take readings of voltmeter and ammeter.
6. Note the temperatures along the length of the specimen rod at 5 different
locations.
7. Note down the inlet & outlet temperatures of cooling water and measure the
flow rate of water.
8. Repeat the experiment for different heat inputs.
OBSERVATION TABLE:

Voltmeter Ammeter Heat Metal rod thermocouple reading Water Volume

Reading Reading input (0C) temp (0C) flow

‘V’ ‘I’ VxI rate of


water,
(volt) (A) (W) Inlet Outlet
V
cc/min
T1 T2 T3 T4 T5 T6 T7 T8 T9 T10 T11

CALCULATION:
Plot the variation of temperature along the length of the rod. From the graph,
obtain dT/dx, which is the slope of the straight line passing through/near to the
points in the graph. Assuming no heat loss, heat conducted through the rod =
heat carried away by the cooling water

-k x A x dT/dx = mf Cp (T10 – T9)

where, ‘k’ = thermal conductivity of metal rod, (W/m-K)


‘A’ = Cross sectional area of metal rod = πd²/4 (m²)
‘d’ = diameter of the specimen = 20 mm
‘Cp’ = Specific heat of water = 4.187 kJ/kg-K
Thus, the thermal conductivity ‘k’ of metal rod can be evaluated.

Mf Cp (T10 − T9)
−𝑘 =
A × 𝑑𝑇/𝑑𝑋
RESULT
The Thermal conductivity of given metal rod is……………………..

PRECAUTIONS:

1. Do not give heater input without the supply of water.


2. Input should be given very slowly.
3. Run the water in the jacket for about 5 min after the experiment.
4. Do not run the equipment if the voltage is below 180V.
5. Check all the electrical connections before running.
6. Before starting and after finishing the experiment the heater
controller should be in off position.
7. Do not attempt to alter the equipment as this may cause damage to
the whole system.

Das könnte Ihnen auch gefallen