87% found this document useful (23 votes)
13K views42 pages

Construction and Application of Heat Sensor 2

This document discusses the construction and application of heat sensors. It begins with introducing heat and temperature, and how temperature sensors are used to monitor and control systems operating within defined temperature ranges. It then discusses how semiconductor devices and LCDs can be damaged by temperature extremes, and how temperature sensing helps improve reliability. The document outlines the objectives of studying heat sensor construction and application, which include examining heat transfer in microprocessors and the components, temperature ranges, and ideal environmental conditions of heat sensors. Finally, it defines key terms like heat, sensor, and microprocessor.

Uploaded by

okereke ebuka
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
87% found this document useful (23 votes)
13K views42 pages

Construction and Application of Heat Sensor 2

This document discusses the construction and application of heat sensors. It begins with introducing heat and temperature, and how temperature sensors are used to monitor and control systems operating within defined temperature ranges. It then discusses how semiconductor devices and LCDs can be damaged by temperature extremes, and how temperature sensing helps improve reliability. The document outlines the objectives of studying heat sensor construction and application, which include examining heat transfer in microprocessors and the components, temperature ranges, and ideal environmental conditions of heat sensors. Finally, it defines key terms like heat, sensor, and microprocessor.

Uploaded by

okereke ebuka
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 42

1

CONSTRUCTION AND APPLICATION OF HEAT SENSOR

CHAPTER ONE

1.0 INTRODUCTION
1.1 BACKGROUND OF STUDY
The word “heat” is made manifest as a result of increase in
temperature. Temperature is most often measured environmental quantities
which correspond to primary sensations-hotness and coldness. This is due to
the fact that most biological, chemical, electronic, mechanical and physical
systems are affected by temperature. In many instances, some processes
perform better within a range of temperatures. Also, certain chemical
reactions, biological processes and even electronic circuits do better within
limited temperature ranges. When the needs to optimize these processes
arose, the systems used for controlling the temperature within defined limits
are then needed. The temperature sensors are often used in providing inputs
to those control systems. However, in the case of too much exposure of some
electronic components to high extreme temperature, there will be an advace
effect on them which can lead to he damage of the components. Though,
some of the components can even be affected and get damaged by low
temperature values.
Semiconductor devices as well as LCDs (Liquid Crystal Displays)
can be affected and get damaged by temperature extreme. As the temperature
threshold gets exceeded, an immediate action should be taken so as to
prolong the lifetime of the system. In these, temperature sensing helps to
improve the reliability as well as the lifetime of the system. Most temperature
monitoring devices are designed to respond to a particular (critical)
temperature level. They are usually incorporated with different kinds of
alarms and light indicator units, which are triggered ON at an unacceptable
temperature level. These temperature monitoring devices work with
temperature sensors normally transducers which generate accurate voltage
output that varies linearly with temperature. They are mainly used for
monitoring industrial machines, electric boilers, ovens and other heat energy
related activities and this can be done by ensuring the temperature sensor and
its leads are at the same temperature as the object to be measured. This
usually involves making a good mechanical and thermal contact.
If the temperature sensor is to be used to measure temperature in
liquid, the sensor can be mounted inside a sealed end metal tube and can then
be dipped into a bath or screwed into a threaded hole in a tank. Temperature
sensors provide inputs to those control systems. When temperature limits are
exceeded, action must be taken to protect the system. In these systems,
temperature sensing helps enhance reliability. In modern electronics, more
temperature measuring techniques are available. Several temperature sensing
techniques are in widespread usage. The most common of these:
Thermocouples, Thermistors and Sensor IC's. These temperature sensors
("transducers"), illustrate a nice variety of performance tradeoffs.
Temperature range, accuracy, repeatability, conformity to a universal curve,
2
size and price are all involved. The temperature sensor being used in this
system is the LM 335 and it has the following features; (1) Directly calibrated
in Kelvin, (2) 1°C initial accuracy available, (3) operates from 400u.A to
5mA, (4) less than the 1H dynamic impedance, (5) easily calibrated and (6)
low cost. This makes it preferable to the thermistor which was used in
previous related research. Although, the thermistor has a wider temperature
range than the LM 335, it suffers from self-heating effects, - usually at higher
temperatures where their resistances are lower -and fragile, which makes it
inappropriate for the research. Apart from the merits of the temperature
sensor, the dual nature of the system extends its application which is not so in
previous systems designed for heat monitoring.
The Complementary Metallic Oxide Semiconductor (CMOS), which
provides reasonable performance, is extensively used in the construction of
this system. In previous research relating to temperature monitoring,
transistor-transistor logic (TTL) integrated circuits were used. The TTL
devices are attributed to high power consumption, limited logic functions,
narrow power supply, low compatibility and high overall cost while CMOS
on the other hand, provides low power consumption, good immunity to
external noise, insensitivity to power supply variations, temperature range
capabilities -48°C to 52CC [1]. In general, the most important characteristics
of CMOS make it the logic of choice. A typical temperature monitoring
device possesses both a temperature sensor and control unit that responds to
the input. It is against this background that the study wishes to examine the
construction and application of heat sensor.

1.2 STATEMENT OF RESEARCH PROBLEM


The heat sensor is very important in the determination of temperature
variation in our environment; it is mostly used in the hospital, heavy
construction industries and petroleum industries. Despite the contribution,
there are some limitation, which include; the inability to be use in all fields
(health and manufacturing purposes). Another problem is the problem of cost
of manufacturing and maintenance. Lastly there have been several studies on
tempreture but not even a single study have been carried on the construction
and application of heat sensor; hence a gap in literature.

1.3 AIMS AND OBJECTIVES OF STUDY


The main aim of the research work is to examine the construction and
application of heat sensor. Other specific objectives of the study include:
1. To examine heat transfers in microprocessors
2. To determine the fundamentals of heat transfer
3. To examine the components of heat transfer sensors
4. To examine the temperature range in various heat sensors
5. To determine the favourable environmental condition for heat sensors

1.4 RESEARCH QUESTIONS


The study came up with research question so as to be able to ascertain the
above stated objectives of the study. The research questions for the study are
3
stated below as follows:
1. How does heat transfer in microprocessors?
2. What are the fundamentals of heat transfer?
3. What are the components of heat transfer sensors?
4. What is the temperature range in various heat sensors?
5. What is the favourable environmental condition for heat sensors?

1.5 ORGANISATION OF STUDY


The chapter one of the research work will contain the introduction, the study
background, the statement of research problem, the aims and objectives of
study, the research questions, and significance of study, the scope of the
study and the definition of terms. The chapter two of the research work will
contain the review of related literature as regard the construction and
application of heat sensor. The chapter three of the research work will contain
the materials and method required for the construction of the heat sensor. The
chapter four of the research work will contain the experimental effect of the
application of the heat sensor. The chapter five will contain the conclusion
and the recommendation of the study.

1.6 SIGNIFICANCE OF STUDY


The study on the construction and application of heat sensor will be of
immense benefit to the entire institution, the department, the students and
other researchers that wishes to carry out similar research on the above stated
topic as the findings of the research work is education people on
fundamentals of heat sensors and also the experimental effect of the
application of the heat sensor. The study finally will contribute to the body of
the existing literature on the construction and application of heat sensor

1.7 SCOPE OF STUDY


The study on construction and application of heat sensor will cover the
fundamentals of heat sensors and also the experimental effect of the
application of the heat sensor.

1.8 DEFINITION OF TERMS


HEAT: Heat is the amount of energy flowing from one body of matter to
another spontaneously due to their temperature difference, or by any means
other than through work or the transfer of matter.
SENSOR: A sensor is a device that detects and responds to some type of
input from the physical environment
MICROPROCESSORS: an integrated circuit that contains all the functions
of a central processing unit of a computer.
4
CHAPTER 2

2.0 REVIEW OF RELATED LITERATURE


2.1 Fundamentals of Heat Sensors
Heat sensors are found in many items, from commonplace items inside any
home to more sophisticated applications. You can find sensors in household
electronics like thermostats or thermometers. You will also find sensors in
things as sophisticated as your personal computer or in a microprocessor. It
is vital for pro- cessors to stay within the temperature range specification to
perform reliably and for the processor to run at its expected speed
performance. In this chapter, we review the fundamental principles of heat
transfer and describe heat transfer in a typical microprocessor package. We
also touch on the principles of heat sensors, including the various sensor
materials, operation and applications in a typical semiconductor industry
environment.

What Is a Heat Sensor?


Sensors are devices that measure a physical or chemical reaction, such as
volume flow or heat flux, through changes in electric resistance or signal
(Kenny 2004). There are many types of sensors—flow, force, pressure,
humidity and motion sensors are just a few. We are focused on one type of
sensor in this book: heat sensors.

2.2 Overview of Heat Sensors


Temperature is the measure of the average kinetic energy of the molecules of
a gas, liquid, or solid. A heat sensor is a device that is specifically used to
measure temperature. In this way, heat sensors are able to give us a
quantifiable way to describe the substance, whether it is an object, the
environment in which an object is placed or the environment in which an
object is distributed. More about how these sensors are applied to
microprocessors are discussed in later chapters.

2.2.1 Types of Heat Sensors


One well-known heat sensor is a mercury or alcohol thermometer. It uses
the volume of mercury or dyed ethanol, which expands when temperature
increases, to measure temperature in a tube with a temperature scale.
Though very well known, mercury and alcohol thermometers are not well
suited to measure temper- ature in a personal computing device or
microprocessor because they tend to be too large for those applications.
Other kinds of heat sensors that can be suited for personal electronics and
microprocessors include thermocouples, resistance thermometers, silicon
sensors and radiation thermometers.

Thermocouples
Thermocouples are sensors composed of two different metals at their sensing
end. A voltage is created when there is a temperature gradient between the hot
sensor element and the cold reference junction. The change in voltage can be
5
reported as a tempera- ture through the Seebeck effect (Love 2007). The
Seebeck effect says that the change in voltage is linearly proportional to the
change in temperature and the two variables are related to each other through
a coefficient that is determined by the materials used in the thermocouple
(Janata 2009). Figure 2.1 depicts the construction of a thermocouple.

Resistance Thermometers
Resistance thermometers are also known as resistance temperature
detectors, or RTDs. They are typically made of a single pure metal (Dames
2008). Each metal has a material property of electrical resistance that is a
function of temperature. The most accurate resistance thermometers are
ones that use metals that have a very linear relationship with temperature,
such as platinum. By using the relation- ship curves between electrical
resistance and temperature, when the resistance of the metal is measured, a
temperature can be calculated (Dames 2008). Figure 2.2 depicts the
construction of one type of resistance thermometer.

Thermistors
A thermistor is a specific type of resistance thermometer. Thermistors are
made of metal wires connected to a ceramic base made of several sintered,
oxide semi- conductors (Janata 2009). Like other resistance thermometers,
the change in

Fig. 2.1 Thermocouple construction. Adapted from Love (2007)

Fig. 2.2 An Example of resistance thermometer construction. Adapted


from Desmarais and Breuer (2001)
6
temperature can be calculated from the change in resistance. But unlike
traditional resistance thermometers, the relationship is not very linear.
Thus, the temperature range in which thermistors can be used is small
compared to traditional resistance thermometers. But thermistors have the
advantages of being small in size, inex- pensive to buy and very sensitive
to temperature changes, so they can be ideal to use in many electronics
applications (Janata 2009).

Silicon Sensors
These sensors are made of silicon, a semiconductor that is used as the base
mate- rial for most electronic microprocessors. The process of
manufacturing these elec- tronic devices is a carefully controlled, high-
volume manufacturing process that includes deposition, doping, and
careful layering of metals, oxides, and insulators (Peterson 1983). By
utilizing this manufacturing process, integrated circuit (IC) sensors can be
created as their own sensor device (Desmarais and Breuer 2001). They can
also be embedded inside microprocessors as diodes (Rotem et al. 2006).
These types of sensors can have their own memory, can have direct output
to meters and can convert signals to temperature readings without extra
equipment (Desmarais and Breuer 2001).

Radiation Thermometers
All substances and objects emit heat radiation when it is at a temperature
higher than absolute zero (0 K− or 273.15 °C). There is a relationship
between temperature and radiation energy emitted that can be used to
calculate the temper- ature of the object surface. Unlike other sensors
discussed above, radiation ther- mometers are primarily used at a distance
from the object of interest and can be used for hard-to-reach objects. An
example of a radiation thermometer is an infra- red camera, which
measures infrared wavelengths that emit from an object.

2.2 Heat Transfer and Microprocessors


Power is an important design feature of microprocessors. It is linked to the
expected silicon performance and also generates the heat that must be
cooled from the part. The goal of microprocessor heat management is to
cool the processor efficiently within a specified temperature to ensure
reliable performance over the lifetime of the part. Thus, it is important to
understand the heat transfer mechanisms in the microprocessor that cool
the processor in order to understand the importance of heat sensors in
microprocessors.

Fundamentals of Heat Transfer


Thermodynamics describes the fundamental behavior of heat and
temperature and includes the three laws of thermodynamics. Heat transfer
goes further and describes the mechanisms of heat exchange and the rate at
which heat flows, giv- ing us a way to calculate heat flow within, to and
from objects or the environment. There are three modes of heat transfer:
7
conduction, convection and radiation.

Conduction
Conduction is the first mode of heat transfer that we will discuss. It is a
prominent mode of heat transfer in electronics cooling. Undergraduate
transfer textbooks typ- ically devote a large portion of the text to this topic
and those books will be able to give a more comprehensive and in-depth
discussion on this mode of heat transfer.

Definition of Conduction: Conduction is the heat transfer through solids.


It can also occur with stagnant fluids. The one-dimensional rate of
conductive heat transfer is determined by Eq. 2.1:
Qconduction kA(Thot − Tcold ) (2.1)
=
L

Fig. 2.3 Heat transfer through an object by conduction where

Qconduction is heat flow, k is the heat conductivity of the material, A is the


cross-sectional area of heat flow, Thot is the temperature of the hot surface,
Tcold is the temperature of the cold surface and L is the length of the
material through which heat is conducting. Figure 2.3 depicts the heat
transfer through a solid mate- rial by conduction. The different variables of
the conduction heat transfer Eq. 2.1 are shown.
The conduction resistance is defined by Eq. 2.2:
Rconduction = L
kA
As shown in Eq. 2.2, in order to minimize the conduction resistance,
conductiv- ity of the material and cross-sectional area of material is
maximized while the through-path (length) of the material is minimized.

Three-Dimensional Conduction: Conduction can also occur in three


dimen- sions if the heat source size is smaller than the conducting material.
If the heat source is smaller than the conducting material, the heat is
concentrated in one spot, causing a hot spot. In Fig. 2.4, heat is being
conducting in the x and y direc- tion as well as the z direction. The
conduction resistance in the x and y direction is also known as the
spreading resistance, which can be solved in idealized boundary conditions
8
by first order equations or more often through computational software.
Figure 2.4 depicts one-dimensional and three-dimensional conduction
through a solid, as well as the transition between the two types of
conduction.

Contact Resistance: Another thing to consider in conduction is contact


resist- ance between two different solid materials. When two materials join
to form a conduction path, there is resistance at the point where the two
materials join. This occurs because the two surfaces are rarely completely
flat and voids of air are pre- sent at the junction of two surfaces. This must
be taken into account when heat solutions are applied to electrical
packages or when an external heat sensor like a thermocouple is attached
to the package. Minimizing contact resistance is a key factor to consider in
electronics packaging and their cooling.

Example 2.1 A rectangular block has an area of 400 mm 2. An engineer


would like to use it to cool his heat source that is producing a total of 50 W
with a specifica- tion of 90 °C. The heat source is placed in a chamber with
an air temperature of

Fig. 2.4 Transitions between one-dimensional and three-dimensional


conduction

65 °C. If the through-length is 10 mm, what is the minimum conductivity of


the material in order to cool the heat source to specification using the block
alone?

Solution
In this example, the mode of heat transfer is conduction through the block.
The given information is annotated in the illustration below to help us solve
this problem.
9

We will start with Eq. 2.1 and replace with the given variables. We want to
solve for conductivity k in units of W m−1 K−1.
Qconduction kA(Thot − Tcold )
=
L

In order to solve the equation with the correct units, the area and length
units must be changed to meters, m. The known values are plugged into
the conduction equation to give:

50W = k(0.0004 m2)(90◦C)


0.01 m

Rearranging to solve for conductivity k:

(50 W)(0.01
m)
k= 2
. Σ
0.0004 m (90
W
k = 50
mK
By re-ordering Eq. 2.1, we solve for k and find it to be 50 W m−1 K−1.

Convection
Convection is the second mode of heat transfer we will discuss. Along
with con- duction, it is typically a large part of electronics cooling in
active, fan-cooled sys- tems. An undergraduate textbook can be consulted
for in-depth information.

Definition of Convection: Convection is the heat transfer from a surface


to a fluid. Some common fluids include air and water. Other fluids such as
alcohol and oil can also be mentioned in cooling electronics. The rate of
convection heat trans- fer is determined by Eq. 2.3 below:
Qconvection = hcA Tambient − Tsurface (2.3)
where Qconvection is heat flow, hc is the convection heat transfer coefficient, A
is the surface area, Tambient is the ambient temperature of the fluid, and
Tsurface is the tem- perature of the surface of the material. Figure 2.5 depicts
1
the airflow profile and temperature profile you would expect through
convection from the surface of the rectangular object. Tsurface is warmer than
Tambient, which corresponds to the lower airflow velocity at the surface.
The convection resistance is described by Eq. 2.4:
Rconvection =
hA c

Fig. 2.5 Convection and its airflow and temperature boundary layers

Fig. 2.6 Air heating: ideal versus real air temperature versus fin temperature

As shown in Eq. 2.4, to minimize convection resistance, the convection


coefficient and surface area should be maximized.

Surface Area and Fins: To increase surface area and the effectiveness of
convection heat transfer, fins are often added to the base of electronic
cooling solutions. The fins are often made of a heatly conductive material
such as alu- minum or copper. Fins can be formed by extrusion from a
large block or they can be formed and stacked on top of the base with
solder.

Air Heating and Pressure Drop: Another consideration is the air heating
along the length of the fin. In an idealized condition, the temperature
difference between the air and the fin along the length of the fin would be
constant. However, in reality, the temperature of the air increases along the
length of the fin and the fin’s ability to remove heat is diminished as the
1
length of the fin is increased. It is not effective to simply make the fins
infinitely long. Figure 2.6 shows the ideal versus real case for air heating.
Additionally, longer fins will have a higher pres- sure drop across the fins
compared to a heat sink with shorter fins. For heat sinks with longer fins,
the total airflow will be reduced and thus cooling will be reduced for a
given fan curve.

Optimizing Fin Performance: The fin stacks themselves have a fin resist-
ance, which is a function of convection coefficient, fin efficiency and fin
surface area. The balance among all three will determine the optimal fin
geometry for a given set of boundary conditions, including cost
considerations and manufacturing abilities.

Radiation
The last mode of heat transfer to discuss is radiation. Typically, this mode
of heat transfer can be more complex than either conduction or convection.
Graduate level classes cover this mode in more detail. In cooling
microprocessors, radiation can play a large part when the primary mode of
cooling is natural convection, with no active fan.

Definition of Radiation: Radiation is the heat transfer between surfaces via


elec- tromagnetic waves. All matter at a nonzero temperature emits
electromagnetic waves, including gases and liquids. However, in many
undergraduate heat transfer textbooks, the focus is only on solids. The rate of
heat transfer is described by Eq. 2.5:
Qradiation = hr A(Tsource − Tsurrounding) (2.5)
where Qradiation is the heat flow, hr is the radiation heat transfer coefficient,
Tsource is the temperature of the source and Tsurrounding is the temperature of
the surround- ings (Bergman et al. 2011). The radiation heat transfer
coefficient can also be described by Eq. 2.6:

hr = εσ (Tsource − 2 (2.6)
sourc + surround
Tsurrounding )(T 2 e T
ing)
where ε is the emissivity of the source and σ (5.67 10−8 W m−2 K−4) is
×
the Stefan-Boltzmann constant (Bergman et al. 2011). The radiation heat
transfer coefficient is similar to the convection heat transfer coefficient,
but the radiation heat transfer coefficient is much more dependent on
temperature as shown by rais- ing the temperature terms in Eq. 2.6 to the
third power (Bergman et al. 2011). The net radiation heat transfer is well
known by Eq. 2.7, where hr in Eq. 2.5 is replaced by Eq. 2.6:
Qradiation = εσ 4 (2.7)
sourc − surround
A(T 4 e T
ing)
Figure 2.7 depicts the heat transfer of an object through radiation. The
variables in Eq. 2.7 are highlighted in the figure.

Components of Radiation Heat Transfer: An ideal radiative surface is


called a blackbody with an emissivity of 1, but in reality, no surfaces are
ideal. Radiative heat transfer is highly dependent on all the bodies
surrounding the surface in
Fig. 2.7 Heat transfer through radiation from an object to its surroundings

Fig. 2.8 Components of radiation heat transfer

question and the temperature and finish of the surfaces. A solid surface can
emit, absorb, reflect, and transmit radiation depending on the material in
question. For example, an opaque material can reflect radiation, whereas in
a semitransparent material, radiation can be transmitted through the
material. Unlike absorption and emission, reflection and transmission do
not affect the total heat energy of the material (Bergman et al. 2011).
Figure 2.8 depicts the various components of heat transfer of an opaque
solid through radiation.
2.3 Heat Transfer in a Microprocessor
As discussed in the previous section, the three modes of heat transfer can
play important roles in the cooling and temperature sensing of
microprocessors. Typically, in active, forced air-cooled systems,
conduction and convection play the largest role. In natural convection
systems or in systems where there is no room to attach a fan, radiation can
play an important role.

Active Cooling
In active cooling systems, fans are attached in the system; they provide
airflow over the microprocessors and can be used to help cool them. In
these cases, the heat engineer can conduct heat from the package out to a
heat sink made of a heatly conductive material such as copper or
aluminum. In cases where more cooling is needed, fins can be added to the
cooling solution. Though actively cooled, it is also important to consider
the conductive path from the package to the motherboard or any other
substrate the microprocessor is attached to. Figure 2.9 depicts a few
examples of heat sinks that can be used to cool microprocessors.

Fig. 2.9 Various heat sink configurations

Natural Convection Based-Cooling


In natural convection cooling solutions, no air movers are present or very
little air is available due to airflow blockages. In these cases, radiation is an
important cool- ing mechanism as well as conduction into the motherboard
and natural convection. Analysis would have to include all the parts in the
system in order to accurately pre- dict heat performance of the
microprocessor or object in question. This analysis is one of the most
complex heat transfer modes to model because all the surfaces, materials,
and properties associated with the object must be specified accurately.

Package Heats
Heat is spreading through conduction within the die and out of the package
via active or natural convection cooling. The package is composed of the
die, which is attached to a substrate made of FR-4 material with embedded
copper layers. The package designer may decide not to cover the die; this
is typically called a bare- die package. Alternatively, a package can include
over-molding made of an epoxy or plastic to protect the die. It can also be
covered with a copper integrated heat spreader (IHS) to help reduce
spreading resistance and increase conduction to the heat sink. If the die is
covered with a spreader, another material is needed between the die and
spreader to fill in small air voids and gaps with conductive mate- rial.
This interface material is typically called the heat interface material. In
Fig. 2.10, three package options (bare die, over-molded, integrated heat
spreader) are shown.
Example 2.3 A heat sink is placed over a package with an integrated heat
spreader. The package itself is attached to a substrate. The heat sink has
metal fins and a metal base. Draw the resistance stack that should be
accounted for in order to accurately find the die temperature (junction).
Make sure to take into account any interfaces that would be present in a
real assembly.

Fig. 2.10 Various package configurations with and without lid protection

Solution
To understand the different interfaces that should be accounted for, a
schematic would be the first place to start. The problem statement gives
the description for what the heat solution looks like. The illustration below
depicts the description given in this example of a heat sink with fins placed
on top of a package with an integrated heat spreader.
The heat sink can be a radial heat sink or one of another shape, but for
sim- plicity, we chose a simple rectangular base with straight fins. In this
problem, it does not matter how tall or long the fins are. The base area and
thickness are also not important. In this problem, the package is a ball-grid
array (BGA) package, attached to the motherboard using solder balls.
Alternatively, the package can be attached to the motherboard using a
land-grid array (LGA) socket.
There are two places where heat interface material would be placed in
this configuration: within the package between the die and integrated heat
spreader (Heat Interface Material 1) and between the integrated heat
spreader and the heat sink base (Heat Interface Material 2). Not only can
power go up from the die into the heat sink, power may also go down
through the substrate. It is also important to know that if power is
dissipated into the substrate, the power will eventually be cooled by T air.
The resistance stack can be created by choosing temperature points
along the path through which heat will be dissipated, as shown in the next
schematic. The die temperature is also referred to as the junction
temperature in many specification sheets and is shown as the junction
temperature in this example. It is important to note that the package
resistance is in reality made of many parts: spreader resistance, heat
interface material 1 resistance, and so forth. However, many specification
sheets simply give an overall resistance target from junction to case, which
should be defined in the specification sheet, rather than describe all the
details.
For this example, the resistance network is shown in the following
illustration. This network reflects the simplification within the package by
combining the spreader and heat interface material 1 resistances into one
resistance and reflects that the power through the package will be cooled
by air.

Need for Temperature Measurement in a


Microprocessor

Each microprocessor needs a mechanism to give feedback to the user and


system on how hot it is getting when it is powered. Without temperature
measurement, the system may not provide enough airflow to cool the part
or the engineer may not know if the heat solution is adequate to keep the
microprocessor within specification.

Hot Spots in the Microprocessor


In an ideal heat situation, the microprocessor would be uniformly heated
when it is powered. However, in reality, the microprocessor is made up of
many different subsections that are designed to do different tasks. Thus, it
is often the case that certain parts of the microprocessor are hotter than
others and hot spots develop on the die where the microprocessor is more
heavily utilized. Where the hot spots are located is dependent on the
application and the layout of the die. The non- uniformity of heating and
its dependence on workload application adds more com- plexity to
understanding the processor temperature and it is not easily modeled
accurately. Sensors are needed to understand where the hottest temperature
is in real-use conditions.

Microprocessor Performance
There are many inputs into determining microprocessor performance,
including voltage, leakage power, and heat design power (TDP). The
inputs are dependent on temperature and subsequently, the performance of
the part, its frequency, is also dependent on temperature. Higher TDP,
higher voltage, and lower leakage power increases frequency. This is the
ideal case when performance is the most important variable in the design.
However, if both temperature increases and reliability metrics are fixed, the
voltage decreases and leakage power increases, leaving less power devoted
to TDP. This is the opposite trend of what is needed for optimal
performance. It is vital to measure the temperature of the microprocessor
accurately to properly set frequency.

Microprocessor Reliability
The silicon microprocessor has an upper temperature limit specification
that is set to prevent immediate microprocessor damage. In addition,
microprocessors have an allowable failure rate and are specified to work
for a certain period of time, usu- ally in the number of years. If reliability
is relaxed and the part can tolerate more failures or a shorter lifetime,
voltage may not need to decrease when temperature increases, which
allows for better performance. To ensure that there is minimal immediate
damage and that the part failure rate is acceptable in its lifetime, the
microprocessor temperature must be accurately understood.
CHAPTER THREE

3.0 MATERIALS AND METHOD


3.1 Sensor Materials
There are many materials that are used for sensors. Metals are typically
used for RTDs and thermocouples. The semiconductor silicon is used to
make micropro- cessors and this is where many heat sensors are built into
the die. The follow- ing is a quick, but by no means an exhaustive,
overview of the materials.
Platinum Sensors
Platinum is the choice material to use for RTDs. The typical temperature
usage range for platinum −RTDs is 250 to 600 °C, but some platinum RTDs
can be used up to 850 °C (Love 2007). Its electrical resistivity has a very
linear relationship

Table 2.1 Physical and heat properties of platinum


Physical property Platinum (Pt)
aHeat conductivity (W m−1 K−1) at 25 71.6
°C
bMelting point (K) 2042
cTemperature coefficient of resistance 3.92 × 10−3
dElectrical resistivity Ω cm at 0 °C 9.6 × 10−6
bCoefficient of linear heat 8.8 × 10−6
expansion (K−1)
eSpecific heat capacity (J g−1 K−1) 0.133
bDensity (g cm−3) 21.45

Sources aHo et al. (1972)


bTouloukian et al. (1975)
c Serway (1990)
dHall (1968)
eWagman et al. (1982)

with temperature over a large range of temperature (Dames 2008) and has a
higher resistivity compared to other metals such as copper and nickel, which
make it ideal for RTDs (The RTD 2014). The temperature coefficient of
resistance (TCR) is 0.0039 K−1 at room temperature, where the TCR is the
change in resistance per unit change in temperature. A higher TCR
represents a more sensitive RTD to tem- perature and a high TCR is ideal.
As a material, it is chemically inert and stable in different kinds of
environments and is not likely to corrode or reduce, making it useful to
measure temperature in many different environments (Dames 2008). Table
2.1 shows a few common heat and electrical properties for platinum.

Thermocouple Materials
Many metals and alloys other than platinum are used in sensors, especially
thermocou- ples. Though RTDs are mostly made of platinum, copper and
nickel RTDs can also be found. In thermocouples, alloys are common materials
(ANSI and IEC Color Codes 2014). Because of the wide range of materials,
the temperature range of thermocouples is much −wider than RTDs:
temperatures can range as low as 270 °C and as high as 2300 °C. Common
non-alloy thermocouple metals are summarized in Table 2.2.
In Table 2.3, a few common thermocouple alloys are summarized. The
alloys in the table are made of copper, nickel, and chromium and they are
found in J, K, T, and E thermocouple types.
Care must be taken in choosing thermo-electric materials for the
environment they are used in: for example, thermocouples containing iron
can be more suscep- tible to oxidization at high temperatures and is
recommended for lower temper- atures (ANSI and IEC Color Codes
2014). Chromium-based thermocouples can also see oxidation or “green
rot” at higher temperatures when exposed to low lev- els of oxygen
(Nicholas and White 2001). Table 2.4 summarizes the recommended and
limiting conditions of a handful of thermocouple types.

Table 2.2 Physical and heat properties of copper, nickel and iron
Physical property Cu Ni Fe
aHeat conductivity (W m−1 K−1) at 401 90.9 80.4
25 °C
bMelting point (K) 1357.6 1728 1811
Temperature coefficient of c3.9 × d5.9 × c5.0 ×
resistance (Ω/Ω °C)
10−3 10−3 10−3
eElectrical resistivity Ω cm at 0 °C 1.545 × 6.23 × 8.7 ×
10 −6 10 −6 10−6
bCoefficient of linear heat 16.5 × 13.4 × 11.8 ×
expansion (K−1) 10−6 10−6 10−6
fSpecific heat capacity (J g−1 K−1) 0.385 0.444 0.449
bDensity (g cm−3) 8.933 8.90 7.87

Sources aHo et al. (1972) bTouloukian et


al. (1975)
cSerway (1990)
dElectrical Conductivity (2014)
eHall (1968)
fWagman et al. (1982)
Table 2.3 Physical and heat properties of copper-nickel and nickel- chromium
alloysa
Physical property Cu-Ni Ni-Cr
Heat conductivity (W m−1 K−1) 20 19.2
(typical)
Melting point (°C) 1260 1427
Temperature coefficient of resistance −0.01 × 10−3 0.4 × 10−3
(Ω/Ω °C)
Electrical resistivity Ω m (typical) 49 × 10−8 70.6 × 10−8
Coefficient of linear heat expansion 18.8 × 10−6 13.1 × 10−6
(K−1)
Density (g cm−3) 8.9 8.73
Source aPhysical Properties of Thermoelement Material (2014)

Table 2.4 Recommended environmental conditions and thermocouple type a

Thermocouple Recommended and limiting conditions


type
E Recommended for inert and oxidizing
environmental conditions. Not recommended for use
in reducing or vacuum environmental conditions
J Recommended for inert, reducing and vacuum
environmental conditions. Not recommended for use
in high temperature, oxidizing environmental
conditions
K Recommended for inert and oxidizing
environmental conditions. Not recommended for use
in a reduced or vacuum environmental conditions
R Recommended for inert or oxidizing environmental
conditions without metal sheath protection
T Recommended for inert, moist, oxidizing and
vacuum environmental conditions

Source aANSI and IEC Color Codes (2014)


Physical property Silicon (Si) Table 2.5 Physical and
aHeat conductivity (W m−1 K−1) at 149 heat properties of
25 °C silicon
bMelting point (K) 1687
cSpecific heat capacity (J g−1 K−1) 0.712
at 300 K
bDensity (g cm−3) 2.42
Sources aHo et al. (1972)
bTouloukian et al. (1975)
cHall (1968)

Silicon Sensors
Silicon is a semiconductor, which at a very basic level is a material that has
an elec- trical resistance between an insulator and conductor (Yacobi
2002). Silicon has a diamond lattice structure and its neutral valence
electron configuration allows silicon to equally share its valence electrons
with other elements (Yacobi 2002). Because of this, silicon can be doped
with other elements near it on a periodic table, such as boron or
phosphorous, to fill its lattice structure with electrons that carry electrical
current, thereby increasing the electric properties of silicon by orders of
magnitude (Berger 2013). Heatly, silicon’s heat conductivity increases as
temperature increases, the opposite of what occurs with most metals
(Yacobi 2002). Most importantly, electronics are often made of silicon for
several key reasons:
• Silica can be purified and turned in single-crystal silicon in a very pure
form by vapor deposition (Habashi 2013).
• It is an inexpensive material (Peterson 1983).
• Its melting point is sufficiently high enough for silicon to be stable
during high volume manufacturing, specifically for high temperature
oxidation, diffusion, and annealing (Yacobi 2002).
• Silicon electronics can be manufactured in batches very precisely (Peterson
1983).
However, one important disadvantage of silicon sensors is the narrow
range of temperature use compared to thermocouples and resistance
thermometers: silicon sensors are generally only good from 50 to 150 °C

(Bakker 2002). Table 2.5 is a summary of some of the electrical and heat
properties of silicon.

Principles of Heat Sensors

There are many considerations to take into account in choosing a sensor to


use. Some questions that frequently come up during that process are:
• What is being measured?
• What cost is acceptable?
• What is the accuracy required for the measurement?
In this section, we will have a quick overview of some factors that play a
part in selecting a sensor for use in the industry, including temperature
range for measure- ment, how the measurement will be made and the
accuracy required.

Objective of Measurement
There can be several reasons why temperature is being measured for a
given process, test, or object. The temperature may have a big effect on the
chemical process that is being monitored. For example, in silicon high
volume manufacturing, the tempera- ture has to be set precisely to control
layer growth and depth. In other cases, the temperature can be the main
output of the test. An example of this case is during the performance test of
a heat heat sink. Temperature can also be measured to ensure reliability
over time. Silicon dies usually have temperature monitors to ensure they
remain under the maximum temperature limit and reliably work over a set
number of years. Whatever the situation, care must be taken to pick a
sensor that can get the job done accurately within the required conditions
without being too costly.

Temperature Range of Measurement


It has been briefly mentioned in the section on sensor materials that each
type of sensor has a certain temperature range of use. For
thermocouples, the general−range of use is between 270 and 1300 °C, but it

can be as high as 2300 °C. For resistance thermometers, the temperature
range of use is between 250 and 600 °C. For silicon sensors, the− useful
temperature range is between 50 and 150 °C. Of the three main sensor
types, silicon has the most restrictive range, while thermocouples can be
applicable in a wide variety of temperature cases.
Within the thermocouple group, different thermocouple types are rated
for dif- ferent temperature ranges. For example, Type J thermocouples,
made with iron and nickel-copper materials, are useful to as high as 1200
°C. However, Type T thermocouples, made with copper and nickel-copper
materials, are only useful to 400 °C (ANSI and IEC Color Codes 2014).
Resistance thermometers also have different temperature ranges
depending on the material used. Platinum is the preferred choice of
material for RTDs and its usable temperature−range is widest, roughly
between 200 and 650 °C. Nickel and copper alloys have the smallest
usable temperature ranges, roughly between 0 and 205 °C, while nickel
and copper have narrower usable temperature ranges between platinum
and nickel/copper alloys (Desmarais and Breuer 2001). Thermistors, which
use semiconductor oxides, have an even narrower range of use, between
−100 and 300 °C (Desmarais and Breuer 2001).
If the conditions in which the thermometers are used reach or exceed
the extremes of the usable temperature ranges, uncertainty can increase
and tempera- ture can deviate from the specifications. Determining the
temperature ranges of the environments in which the sensors are being
used during testing, processing and over their lifetimes will prevent these
inaccuracies in measurement.

Environmental Conditions for the Heat Sensors


Each type of sensor can also have limitations in certain environments.
Material type will play a large part in this consideration. For example, iron
and nickel-iron alloys can oxidize above 535 °C and corrode (Desmarais
and Breuer 2001). These materials should be avoided under these
conditions. Platinum is a very inert mate- rial and can be used in high
temperatures and moist environments without any problem (Dames 2008).
Each sensor with limiting materials can also be protected with special
sheaths and materials under extreme environments, but this may change
the size and cost of the sensor. In addition, some sensors, like thermocou-
ples, would withstand higher shock and vibration conditions than
thermistors or other resistance thermometers (Desmarais and Breuer 2001).
In general, all envi- ronmental conditions—temperature, humidity,
exposure to moisture, exposure to shock and vibration—must all be taken
into consideration in choosing a sensor.

Cost
Cost can be an important factor in choosing a sensor. Platinum is an
expensive material and because of the resistance of the material, it
typically will result in a long sensor element (The RTD 2014). Copper,
nickel, and lead are very low cost materials that can make them more
appealing despite their material limita- tions, such as a narrower use-
temperature range. Because silicon sensors can be produced in batches,
their cost can also be quite low compared to resistance ther- mometers, but
their temperature range is much narrower in comparison. In addi- tion, the
cost of the sensor can increase as the accuracy required increases. To
approximately estimate the cost of a thermometer, including the costs of
the mate- rial, manufacturing, calibration and metrology equipment
needed, the following calculation can be used (Nicholas and White 2001):
USD 100oC
Cost =
Accuracy Required

Using Eq. 2.8, a thermometer requiring an accuracy of 0.01 °C will be 500


times more expensive than a thermometer requiring an accuracy of 5 °C, with
the former costing an estimated USD 10,000. Thus, it is very important to not
only pick the correct mate- rial, but the correct level of accuracy required to
limit the cost of the thermometer.

Accuracy
Accuracy is the ability of the thermometer to exactly hit a specified
temperature. This can be considered a± qualitative description. In contrast,
uncertainty is the
± range of expected error between the actual and ideal
temperatures (Kenny 2004). Thermometer A can have an uncertainty of 0.55
°C and thermometer B can have an uncertainty of 1.3 °C quantitatively, while
thermometer A can be described qualitatively as being more accurate than
thermometer B.

Resistance Thermometers and Accuracy: Resistance thermometers can


be accurate to 0.15 °C. Most platinum RTDs follow the International
Electrotechnical Commission (IEC) standard 60751, which has equivalent
standards published by the Deutsches Institut für Normung (DIN) and

American Society for Testing and Materials (ASTM) (Dames 2008). The
IEC accuracy standard is divided between Class A and Class B accuracy,
where Class A is stricter and covers a smaller tem- perature range between
200 and 650 °C versus Class B, which has a range up to 850 °C. The
formulations for temperature uncertainty are shown in Eqs. 2.9 and
2.10 (Dames 2008):
Class A: Uncertainty = ±(0.15 + 0.002|T |) ◦C (2.9)
Class B: Uncertainty = ±(0.3 + 0.005|T |) ◦C (2.10)
Table 2.6 shows the expected uncertainties calculated over a range of
temperatures using Eqs. 2.9 and 2.10. Class B RTDs are easily found and
bought and they will be less expensive than Class A RTDs. When buying
RTDs, the class and required uncertainties should be specified.

Thermocouples and Accuracy: Thermocouples are generally less


accurate than platinum RTDs. Thermocouples are classified according to
type (J, K, and so on), where each type is distinguished based on the
different metal combinations used to make the thermocouple.
Manufacturers may follow the ASTM E230 spec- ification listed in Table
2.7 for several thermocouple types. Some suppliers also have
thermocouples that follow tolerances based on the IEC standard, which
may have different tolerances from the ASTM E230 standard.
Even though manufacturers sell thermocouples of various types, it is
always best to check them for accuracy in the environment they will be used
in. Thermocouples do not always arrive within specification. Also note that
care must be taken to specify which standard, type and temperature range is
needed when ordering from suppliers.

Silicon Sensors and Accuracy: Silicon sensor accuracy is dependent on


the process tolerances in manufacturing the silicon sensor. Uncertainties
are also introduced during the conversion from analog to digital
temperature as well as using power supplies with their own uncertainties
(Sharifi et al. 2008). Typically,

Table 2.6 IEC accuracy Class A and Class B


standard—uncertainties of

Uncertainty at Class A Class B


temperature (°C) (°C) (°C)
−200 ±0.55 ±1.3
0 ±0.15 ±0.3
200 ±0.55 ±1.3
400 ±0.95 ±2.3
600 ±1.35 ±3.3
800 N/A ±4.3

Table 2.7 ASTM E230


standard: standard limitsa
Type Temperature Tolerance, whichever
range (°C) is greater
E 0–900 1.7 °C or 0.5 %
J 0–750 2.2 °C or 0.75 %
K 0–1250 2.2 °C or 0.75 %
R, S 0–750 1.5 °C or 0.25 %
T 0–350 1.0 °C or 0.75 %

Source aANSI and IEC Color Codes (2014)

the accuracy of silicon sensors is between 0.5 and 3 °C (Bakker 2002),


though higher sensor errors have occurred in experience.

Response Time
In addition to cost, accuracy, and useful temperature range, the rate for the
meas- ured object or substance to change temperature should also be taken
in account. For example, if the rate of temperature change for the
substance in question is very fast, then the thermometer choice should also
have a quick response. If the ther- mometer response is too slow, then it
may inaccurately report the temperature at a given time and may allow the
temperature to exceed a specified limit. To get an accurate temperature
reading, the thermometer should be in equilibrium with the system being
measured. After heat is added, it will have to move by conduction or
convection from the source closer to the thermometer and then be
conducted into the thermometer—process steps that will all take time.
Equation 2.11 shows the time it will take the thermometer to respond with
system temperature change as a time constant, when the mode of heat
transfer is conduc- tion (Nicholas and White 2001):

τ = C/-∑KA (2.11)
L
where C is the heat capacity of the fluid or object in J K −1 and the
denominator is the inverse of the conduction resistance in Eq. 2.2. The time
constant can also be written in terms of convection rather than conduction by
using the inverse of the convection resistance in Eq. 2.4. Equation 2.12 would
result (Tomsen 1998):
τ = C/(hcA) (2.12)

The heat capacity is a material property that describes the amount of heat
required to change the temperature of the object by one degree. It is found
in units of J K−1, which shows the relationship between heat, power
and temperature. A larger heat capacity means that it will take more heat to
change the temperature of a substance. Thus, generally, the higher the heat
capacity of the object, the longer it will take for the temperature to change.
Error between the thermometer and sys- tem will decrease exponentially
with the time constant (Nicholas and White 2001).

Heat capacity is also function of mass. Oftentimes, the heat capacity is


described in terms of a constant mass and the resulting specific heat
capacity would be in units of J kg −1 K−1. If the thermometer is large, the
thermometer will have a longer response time. The thermometer will
require more energy to change by a degree. Bigger thermometers like
platinum resistance thermometers can have problems if quick response
times are needed. However, small sensors like thermo- couples will have
quicker response times (Desmarais and Breuer 2001).

Calibration
The sensor’s accuracy can be increased from the production or supplier
specifi- cations through calibration before test. Calibration can be done in
the laboratory before test, or it can be completed by the manufacturer,
usually at a cost. There are two basic ways that sensors can be calibrated,
using fixed points or through bath calibration.

Calibration Using Fixed Points: In this context, fixed points refer to


tempera- ture points that are associated with thermodynamic properties of
a pure substance, such as the triple point or melting point of a substance
(Nicholas and White 2001). These points consistently occur at the same
temperature and are highly reproduc- ible, making them one of the most
accurate ways to calibrate a thermometer. The International Temperature
Scale of 1990 (ITS-90) is the adopted standard tem- perature scale that
defines temperature in relation−to fixed points. On the ITS-90 scale, the
fixed points range from 270 to 1084 °C in the original scale (Preston-
Thomas 1990), with secondary points added for temperatures up to 3414
°C (Bedford et al. 1996). One of the most important points widely used is
the triple point of water, which occurs at 0.01 °C.

Bath Calibration: In bath calibration, the thermometer in question is cali-


brated against a standard or reference thermometer. Both are placed in the
same environment, such as a bake chamber, a bath, or vacuum chamber
(Dames 2008). Like in calibration using fixed points, it is important for
both the thermometer to be calibrated and reference thermometer to be
correctly attached or immersed in the chamber or bath. In addition, the
reference thermometer must also be very accurate. In this type of
calibration, many temperature points can be tested quickly and calibration
can be completed for a batch of thermometers for a range of test
temperature (Nicholas and White 2001). This kind of calibration can be
completed for resistance thermometers, thermocouples, or silicon sensors.
Direct Versus Indirect Measurement
When a measurement is needed, another consideration to take into
account is where the sensor is located. If the object is very far away,
attaching a sensor directly on or in the object may be impossible.
However, not attaching a sensor on an object of interest can introduce
measurement errors. We briefly discuss direct and indirect measurements
below, with a few examples of each.

Direct Measurement
In direct measurement, the sensor has direct contact with the object or
substance that is being measured. Some examples of direct measurement
include:
• Attaching a thermocouple to the center of an integrated heat spreader of
a microprocessor to measure the case temperature of the package.
• Using a sheathed, platinum resistance thermometer to measure the bath
temper-
ature of water.
• Incorporating an on-die silicon diode onto a microprocessor to measure
the tem- perature of the active area of the silicon.
• Placing thermocouples upstream and downstream of test setup, in the flow
stream
of the test, in order to measure the inlet and outlet air temperature of the
test.
Direct measurement is often the most accurate way to take a temperature
of an object. Even in direct measurement, making sure the correct location
and correct number of sensors used is still important. For example, if the
bath of water is tak- ing up the volume of an Olympic-size pool, using one
resistance thermometer may not capture temperature variation of different
points in the pool of water.

Indirect Measurement
In indirect measurement, contact is not made with the object or substance
of meas- urement. Radiation thermometers make almost exclusively
indirect measurements with no contact with the object. However,
thermocouple and resistance thermom- eters can also be making an indirect
measurement if they are not placed exactly at the location of interest. Some
examples of indirection measurement include:
• Using a radiation thermometer to measure the temperature of an object
at a far distance.
• Placing an on-die diode that is on the die but not directly at the hot spot
of the
processor during workloads.
• Using an infrared camera to measure the temperature contours of a
powered-on processor.

Temperature Scales
Temperature scales that are linked to heat sensitivity were not always in
exist- ence and it has taken many years to standardize the scales (Biró
2011). The first thermometers, where scales were included next to a
measuring tube, are associated with several people, including Galileo and
Ferdinand II of Tuscany (Biró 2011). For most scientists and engineers
today, the most well-known scales are the Celsius, Fahrenheit, and Kelvin
scale. In this section, we discuss these relevant scales. We also briefly
discuss the more recent International Temperature Scale of 1990.

Fahrenheit
The Fahrenheit scale was named after Daniel Fahrenheit, who created this
scale in 1724 (Biró 2011). This scale, like other scales of the time, is
based on using fixed points to determine extremes of the scales and
divided by an easy to remem- ber number of steps (Nicholas and White
2001). Fahrenheit’s scale is sometimes called the 96-based system,
because he divided his scale by 96 parts. At the high extreme, he used the
human body temperature of 96 °F as one of his fixed points and at the low
extreme, he used the melting point of salty ice to be 0 °F (Biró 2011). This
scale is often related to the Celsius scale, by the Eq. 2.13 below:
◦F = 32 + Σ Σ ◦C
5
This equation is not always convenient. But most remember the relation by
know-
ing that 32 °F is associated with the freezing point of water at 0 °C and that
212 °F is associated with the boiling point of water at 100 °C.

Celsius
The Celsius scale is also a well-known and widely used temperature scale.
It is named after Anders Celsius, who created the scale in 1742 (Biró
2011). Like the Fahrenheit scale, it is based on using fixed points at the
ends of the scale but instead of dividing by 96 parts, Celsius divided by
100 parts. For fixed points, he used the boiling temperature of water at the
high end of the scale and the freezing temperature of water at the other end
of his scale. Equation 2.11 can be rewritten to obtain degrees Celsius from
degrees Fahrenheit by Eq. 2.14:
◦C = Σ Σ × .◦F − 32Σ
9

Kelvin
The Kelvin scale was created to describe the absolute zero temperature
point. Absolute zero is described as the point where there is no
thermodynamic motion and where the heat energy is zero (Biró 2011).
Through experiments, absolute
− zero was determined to be 273.15 °C.
Like the Celsius and Fahrenheit scales, the Kelvin scale is fixed at two
points: 0 K is set to absolute zero and 273.15 K is

set to the triple point of water. Because of how absolute zero and 0 K are
defined, degrees Celsius and kelvin are often related by Eq. 2.15:
K = ◦C + 273.15 (2.15)
Unlike degrees Celsius and degrees Fahrenheit, kelvin is not reported as
degrees Kelvin. Instead, it is just report as K, a unit of temperature
(Nicholas and White 2001).
ITS Scale
As mentioned earlier under calibration with fixed points, the International
Temperature Scale of 1990 (ITS-90) is an adopted standard temperature
scale that defines temperature in relation to fixed points. On the ITS-90
scale, the fixed− points are from 270 to 1084 °C in the original scale
(Preston-Thomas 1990). An addition with quality secondary points was
added for temperatures up to 3414 °C (Bedford et al. 1996). The fixed
points, which occur consistently at the same tem- perature and are highly
reproducible, are chosen from thermodynamic points— boiling point,
melting point, or triple point, to name a few—of substances such as water,

hydrogen, copper, and more. On the original − the lowest fixed point
scale,
was the vapor pressure point of helium at 270.15 to 268.15 °C and the

highest point was the melting point of copper at 1084.62 °C. With the
secondary reference points, the lowest fixed point is now zinc at 272.3 °C
and the highest point is the melting point of tungsten at 3414 °C. The ITS
scale is often used to provide a ref- erence calibration point.
Sensing Noise
Sensors are manufactured with specified accuracy ranges, which can
further be calibrated by the manufacturer or user to gain even more
accuracy. Despite the knowledge and experience on obtaining accurate
sensors, readings can still have inaccuracies due to noise. Noise can be
caused by process, metrology, or it can be intrinsic to the sensor. This
section briefly discusses those three sources of noise.

Process Noise
Many manufacturing processes are now very controlled and products from
the manufacturing line are consistent and are mass copies of one another.
Despite this control, it is impossible to hit the same values, such as
thickness or diameter, every single time for every single product off the
line. Manufacturing processes can have some standard deviation from the
mean or set point, which the end user can see as ranges of accuracy in
sensors or as sensors that are not made to an exact specified length or
diameter. A good example of process noise occurs in silicon
manufacturing. This vari- ation will be present in diodes that are embedded
in the microprocessor or in stand-alone silicon sensors. To make silicon
sensors and microprocessors, the manufacturing process involves growing
and laying down several layers of silicon, oxides, and metals, each with its
own thickness set point. If the layer is supposed to be 10 microns thick, it
likely can vary between 10, 10.1, or 9.9 μ, depending on the tolerances
allowed.
Sensors can be rated to be accurate to± 5 °C, but a specific sensor could
be accurate to 1 °C or perhaps 4 °C. When sensors are delivered from the
manufac- turing plant, process noise will be present. Calibration or tests
versus temperature references (that is, a well-stirred water bath) can be
used to understand how pro- cess noise may affect the performance of the
sensors and to reduce any errors.
Metrology Noise
In addition to process noise, metrology noise is introduced through
calibration and measurement equipment, such as chambers, power supplies
and data acquisition machines, to name a few sources.

During Calibration: In order to calibrate sensors, the user or


manufacturer must use reference points by calibrating to fixed points or
reference thermometers. If the reference thermometer is not exactly
accurate itself or if it has drifted over time, then the sensor being calibrated
to it can see this error in accuracy. When baths or chambers are being used,
the temperature distribution within the bath and chamber can introduce
errors to the sensors being calibrated in them. If the tem- perature at the
edge of the chamber is 0.2 °C warmer or cooler than the center, sensors
near the edge would not be calibrated to the same temperature as those
located in the center and a 0.2 °C error is introduced. It is best to make sure
the bath and chamber is well-stirred and that the temperature distribution is
as even as possible during calibration. It is also advisable for the reference
sensors to be cali- brated on a periodic basis.

During Measurement: During measurement, power supplies, data


acquisition machines, temperature readers, voltmeters, multi-meters and
similar equipment can be used in providing power, in recording data and in
reading out data. These electronics will have specific tolerance ranges and
accuracy limits separate from those of the sensors. Calibration may also be
required for the electronics used in gathering measurements from the
sensors. If, for example, the power supply is deviating slightly from the
specification and is providing more voltage than is being reported, the
sensor can read a higher temperature than it would at the cor- rect voltage.
It is advisable to use calibrated electronics with the sensors and to
compensate for any errors introduced if needed.

CHAPTER FOUR

4.0 DATA PRESENTATION DATA ANALYSIS AND


INTERPRETATION
4.0 Sensor Intrinsic Noise: Material and Heat Effects
In all heat sensors, there is noise that is associated with how the sensors
work, with the types and quantity of materials used, and with the long
periods of heat exposure. This kind of noise can be unavoidable, but can be
reduced to minimize its effects. Ignoring or being unaware of these noise
factors may introduce large errors in measurement.

Self-Heating: Self-heating errors occur when the temperature reading in


the sensing element is increased due to heat caused by the sensing element
itself. For resistance thermometers, the current flowing through the wires
causes self-heating. The increase in temperature is described by Eq. 2.16,
R × I2
OTemperatureself −heat =
D
where R is the resistance of the element, I is the current flowing through
the ele- ment, and D is a constant that describes the dissipation between the
element and the fluid or solid being measured (Nicholas and White 2001).
The relationship between the self-heat error and current tells us that even a
small current can produce a large error because current is raised to the
second power. To reduce self-heating effects, the current should be as small
as possible while balancing the required accuracy of the temperature
reading (Dames 2008). Self-heating effects can also be decreased by
maximizing the constant D as much as possible—in order to do this, the
con- tact between the sensor element and fluid or solid should be very
good, with no gaps and no weak attachment methods. Despite the efforts to
decrease self-heating errors, in some cases it will be impossible to avoid
them—for example, in cases where the sensor is measuring the
temperature of very stagnant air (Nicholas and White 2001). Stagnant air
would not provide very effective cooling and it is likely that the whole
sensor setup would self-heat in this environment.

Thermoelectric Effects: Thermoelectric errors occur when more than one


type of metal is used in the construction of the sensing element and
associated body and wiring (Dames 2008). Other areas where different
metal wires can be introduced include metrology electronics, such as data
acquisition machines, and power supplies. When more than one metal is
used, the different metals can pro- duce a Seebeck voltage that would lead
to a temperature gradient along the sens- ing element and associated wires
(Dames 2008). For resistance thermometers, this would lead to different
voltage readings along the wire and introduce error to the temperature
reading of the sensing element. The Seebeck effect is vital in order for
thermocouples to work properly, but introductions of different metal types
in the construction of thermocouple extension wiring can produce
temperature gra- dients in the wires in a similar manner to what occurs in
resistance thermometers (Nicholas and White 2001). In both sensor types,
the degree of error is dependent on the Seebeck coefficient of the metals
involved. The temperature gradient in the wiring can be as much as 10 °C
and consequently, this temperature gradient can lead to inaccurate readings
of 1 °C or more for resistance thermometers and ther- mocouples (Dames
2008); Nicholas and White 2001).

Other Heat Effects: The long-term exposure to heat has the effect of
elas- tically deforming the wires or leads of the sensor. With deformation,
the metal can stretch and contract through several cycles, introducing
errors as a result of differ- ent rates of expansion and contraction and the
subsequent change in the resistance of the wire (Nicholas and White 2001).
In addition, other causes of heat gra- dients than the two previously
discussed (self-heating and thermoelectric effects) will produce noise. For
example, heat effects that lead to larger heat gra- dients can be introduced
from inadequate connection between sensing wires and leads or extension
wires (Nicholas and White 2001).
Sensor Intrinsic Noise: Lead Wire Resistance
For thermocouples, if the correct instrumentation is used with little current
through the circuit, there should be negligible resistance in the wires
(Nicholas and White 2001). Lead wire resistances have a larger effect on
resistance ther- mometers, because current needs to flow through the
sensors for RTDS to work. Resistance thermometers have several wire
configurations with different numbers of wires: two-wire, three-wire, or
four-wire. The circuit they are based on is com- monly called the
Wheatstone bridge (Love 2007). One of the resistors in the cir- cuit is the
resistance thermometer of interest. The other resistors are defined to
complete the circuit in equilibrium at a reference temperature (Love 2007).

Two-Wire Resistance Thermometers: A two-wire configuration is a


common and inexpensive configuration for resistance thermometers.
However, it is also the configuration that is most susceptible to error.
Figure 2.11 depicts the configura- tion of a simple two-wire bridge.

Fig. 2.11 RTD—simple two-wire resistance network. Adapted from Love


(2007)
Fig. 2.12 RTD—two-wire resistance network with long leads. Adapted from
Love (2007)

The bridge works well if the leads to the resistance thermometer are
negligible and there is no measurable resistance in those leads. Error is
introduced when the leads to the thermometer are very long, so that the
lead resistances are not equal or small (Love 2007). Figure 2.12 depicts
where the extra resistors would be intro- duced in series with the resistance
thermometer. The extra resistances would cause inaccurate readings for the
resistance thermometer.

Three-Wire Resistance Thermometers: A three-wire resistance


thermometer features three leads from the resistance thermometer, as
depicted in Fig. 2.13. It is more accurate but also more costly than a two-
wire circuit. RLEAD1 and RLEAD3 will cancel out in this configuration when
the circuit is in equilibrium (Love 2007). RLEAD2 is the resistance between
the voltmeter and the lead across the cir- cuit. The network should have
very little current through RLEAD2 when the circuit is balanced so its
resistance should be very small (Love 2007). In real conditions, there
could be some error due to the lead resistances because no two resistors are
exactly equal, but the error in a three-wire circuit is much less than using a
two- wire circuit.

Four-Wire Resistance Thermometers: The most accurate circuit


configura- tion for a resistance thermometer is the four-wire circuit, as
depicted in Fig. 2.14. It is the most expensive because it uses more
material than a two-wire or three- wire circuit. In this configuration, the
lead resistances carry constant current in the outer loop and another inner
loop with two other resistors measure the resistance across the
thermometer directly (Love 2007).
Fig. 2.13 RTD—three-wire resistance network. Adapted from Love (2007)

Fig. 2.14 RTD—four-wire resistance network. Adapted from Love (2007)

Sensor Reliability
Heat sensors will be exposed to temperature cycling and high temperatures
throughout their use. Inevitably, this will cause degradation in sensor
accuracy and complete breakdown of the sensor as it reaches the end of its
life. In addi- tion, there are other sources that can cause complete sensor
error or breakage well before the end of its expected lifetime. There are
techniques and sensor configura- tions that can prevent catastrophic failure
and extend its life, but sensors are not meant to last forever. This section
discusses a few sources that contribute to unreli- able accuracy and causes
of sensor failure.

Shock and Vibration


Two causes of sensor failure and sensor reading drift are shock and
vibration dur- ing packaging and shipping from the supplier to the
calibration laboratory and ultimately to the use location. Sensors can also
experience shock and vibration effects during attachment and during use
while measuring the fluid or object tem- perature. Shock and vibration can
cause total failure by breaking the sensor wires completely and cause slow
deterioration and accuracy drift if the effects are expe- rienced over time,
for example, as a sensor that is attached to vibrating machinery would
(Nicholas and White 2001). Shock and vibration can also cause the sensor
wires to kink, bend or change shape, leading to changes in electrical
resistance (Nicholas and White 2001). In addition, any metal strengthening
due to plastic deformation caused by shock and vibration can change the
electrical resistance of the wire (Nicholas and White 2001). To prevent
shock and vibration effects, the sensors should be adequately packaged
and carefully handled during shipping. Insulation can also be used to
dampen the shock and vibration effects during ship- ping as well as use.

Hysteresis
Hysteresis is a condition where, in addition to the current sensor
environment, the previous sensor environment also affects the current
sensor readings (Kenny 2004). Hysteresis becomes apparent for
temperature sensors during temperature cycling (Nicholas and White
2001). The metal wires used in sensors will stretch and contract as the
temperature cycles between temperatures. The different rates of expansion
and contraction will affect the next sensor reading and result in errone- ous
readings. As an example, let us take a sensor that is used to measure a fluid
at a high temperature of 400 °C and is then used to measure fluid
temperature as its temperature decreases. The sensor at 375 °C will likely
not have had enough time to fully contract and will still see expansion
effects from being exposed to 400 °C. The reading at 375 °C will
subsequently have errors. As the number of cycles increase, the error also
increases. Hysteresis effects can be reduced by allowing the sensor to
contract back and reach equilibrium at a specific measurement point
(Nicholas and White 2001). However, after many heat expansions and
contrac- tions, the sensor will reach a point where deformation will no
longer be elastic and the sensor will deform permanently. At that point, the
sensor will be irreversibly damaged and its resistance will be permanently
changed.

Chemical Environment
Because many heat sensors and sensor peripherals are made of metals or
alloys, the chemical environment to which the sensors are exposed
becomes a very important factor for long-term reliability. Several different
of types of thermocou- ples were described in an earlier section of this
chapter, including ones made of iron, nickel, platinum, nickel-iron, copper-
nickel, and other alloys. Each metal has a different reaction to moisture,
vacuums, oxidizing, and reducing environments. For example, iron-based
thermocouples may not be suitable at high-temperature, oxidizing
environments because iron is easily oxidized (ANSI and IEC Color Codes
2014). In addition to the wires themselves, thermocouple wire insulators
are subject to their own behaviors in different chemical environments. For
exam- ple, many materials used as insulators for thermocouples break
down in reducing environments, leaving the bare wires exposed to high
temperatures (Nicholas and White 2001). Chemical environment effects
lead to electrical resistive changes and to changes to the Seebeck
coefficient of the metal materials (Nicholas and White 2001). This results
in lower accuracies and can ultimately break the sensor. To avoid negative
environment effects, heat sensor type must be carefully chosen.

General Heat Effects


Because heat sensors are used to measure temperature, sometimes at very
high temperature, heat sensors are naturally going to suffer from heat
effects. The heat effects are often compounded because of sensor exposure
to unfit chemical environments or to vibration at the same time, as was
already described earlier in this section. High temperatures will cause the
metals to generally change shape, internal electron structure, and even
external physical dimensions, which all change the sensor’s resistance
properties (Nicholas and White 2001). As tem- perature increases, the
metals within one sensor may have different heat expan- sion rates,
causing work hardening (Dames 2008). High temperatures can also lead to
metal migration in the wires, leading to metal contamination and increas-
ing electrical resistance (Nicholas and White 2001). Because of these
effects, it is important for the sensors to have proper insulation and for
them to be used in the proper temperature environments. If sensors are not
chosen carefully for measuring capability, the sensors may fail or result in
erroneous measurements. Additionally, sensors should go through
reliability tests to ensure that the heat effects are understood before
permanent installation.

Heat Sensors in Handheld Devices


Smartphones and tablets have become very popular devices and they are
present in many households in replacement of traditional laptops or
landline phones. Because of widespread smartphone and tablet use, there
has been a push to use the handheld devices to do more than their
traditional functions of making phone calls, text messaging, or browsing
the Internet. The ability of smartphones and tab- lets to connect to the
Internet or to a wireless communication network also makes them good
candidates for use as devices that collect information via sensors and share
them with other applications or the Internet for consumption (Fujinami et
al. 2013). In handheld devices today, there are many embedded sensors
that are com- monly used. Accelerometers are used to detect which
direction the handheld device is being held and they are also used in
connection with applications that track running speed. There are also
sensors that detect sound, which can be used to detect surrounding noise
like music for music-recognition applications and light, which is frequently
used today to automatically adjust screen brightness to accommodate
changing dark or bright ambient lighting. In addition, there are GPS,
proximity sensors, gyroscopes, and compasses embedded into smartphones
that enhance location-based applications (Lane et al. 2010).
Within the handheld processor and attached to the board of the phone,
heat sensors can be used in traditional roles of monitoring the temperature
of the pro- cessor and electrical hardware. These sensors are typically
built-in silicon diodes or thermistors. However, despite the multi-
functionality of smartphones and tab- lets, there are not as many instances
where embedded sensors are being commer- cially used to monitor
environmental readings such as temperature, humidity, or UV-radiation
(Fujinami et al. 2013). Studies and development opportunities exist to
expand the use of handheld devices to encompass heat sensors. For exam-
ple, there is a published proposal by Fujinami et al. to use smartphones as
devices that can monitor temperature for heat stroke prevention by using a
silicon sensor and Android†-based software (Fujinami et al. 2013).
However, because handheld devices can be placed in several different
positions, the inaccuracy of the temper- ature readings to the intended
measurements can be large. For example, a hand- held device placed in a
bag will be more inaccurate when compared with a device hanging from a
person’s neck in measuring the ambient environment temperature
(Fujinami et al. 2013). The complication of inconsistent handheld device
place- ment would also make it difficult to use handheld devices to
monitor body or skin temperature for medicinal purposes. Some other
considerations for smartphone and tablet temperature monitoring to be
effective include development of a robust and standard sensing
methodology with handheld devices and agreement on where data is being
stored and how it is transmitted (Lane et al. 2010). In addition, if heat
sensors are to be used to monitor health and ambient conditions
uninterrupted, CPU bandwidth, memory use, and battery consumption with
other applications in the handheld device should be considered and will
need be balanced (Lane et al. 2010).

Heat Sensors in Remote Applications


As sensor technology becomes more sophisticated, smaller, and cheaper,
there have been research and proposals to use sensors wirelessly to
measure an object or fluid from afar or to embed them into textiles to
provide on-body sensing. In this section, we briefly discuss heat sensors as
wireless sensors and in wearables. Like the previous section on heat
sensors in handheld devices, the discussion is not exhaustive but highlights
the emergence and development of how heat sen- sors can be used in the
future.

Smart Sensors
At the basic level, wireless sensors and sensors for wearables can be
consid- ered smart sensors. Traditional sensors are hard-wired,
specialized to one task, localized, and require support equipment to
transform the electrical input to the intended output like temperature
(Mekid et al. 2010). Smart sensors can have the ability to collect various
types of sensor data in one unit or die, like temperature, pressure, and
humidity (Roozeboom et al. 2013). Smart sensors can also integrate
various other functions to one sensor unit, including the ability to
transform the input to a consumable output within the wireless unit,
compensate for expected errors, and communicate and transmit data
(Mekid et al. 2010). For temperature sensors, they can be made “smart” by
integrating the analog-to-digital converter (ADC) with the temperature
sensor (Bakker 2002). Temperature sensors are also likely to be silicon-
based, like doped-Si resistance thermometers or bipolar tran- sistor-based
heat sensors (Roozeboom et al. 2013).

Wireless Sensors
The temperature sensor can be one node in a network of many temperature
sens- ing or other sensing nodes (Mukhopadhyay 2013). For wireless
sensors, commu- nication is typically transmitted through radio frequency
signals. Currently, three of the more popular signals to use are Wi-Fi†,
Bluetooth†, and Zigbee† (Mekid et al. 2010). The radio frequency should
be chosen for the application so that there is minimal interference from
other radio sources, like microwaves or television, and so that the wireless
sensor can operate no matter the location (Mukhopadhyay 2013).
Additionally, the cost and power-bandwidth of each type of signal should be
considered. For example, Wi-Fi has large range and more data bandwidth
capability, but it also consumes large amounts of power compared to using
Zigbee signals (Mekid et al. 2010). There is an increasing interest in sensors
to wire- lessly monitor for applications such as environmental efficiency,
transportation, in smart homes and in health care (Mukhopadhyay 2013). It
is easy to envision the latter two areas as examples where temperature
sensors can easily be applicable. Temperature sensors may be used to
coordinate room temperatures within the home for increased energy
efficiency or comfort. Temperature sensors can also be useful for health-
care monitoring to measure body and skin temperature.

Smart Sensors for Wearables


For health-care monitoring, smart sensors can be embedded in a wearable,
which can be defined as anything from traditional clothing or small
electronics like wrist- watches, wristbands, glasses or chest-bands (Anliker et
al. 2004). There have been several attempts to use temperature sensors to
monitor body temperature, such as in the advanced care and alert portable
telemedical monitor (AMON) project (Anliker et al. 2004), in the
Bioharness† monitoring system (Johnstone et al. 2012), and in protective
equipment to protect firefighters (Talavera et al. 2012). Even though there is
high interest in monitoring temperature for health care, the AMON and
Bioharness monitoring systems have shown inaccuracies in using wearable
temperature sen- sors. For the AMON system, the wristwatch approach does
not provide an adequate means of predicting body temperature (Anliker et al.
2004). For the Bioharness chest-belt, readings from an infrared camera have
low correlation to actual body temperature measurement by a calibrated
thermistor attached to the skin (Johnstone et al. 2012). While these studies
are not exhaustive, it does showcase some attempts at using temperature
sensors in wearables and is also indicative of needed improve- ments in the
future to provide accurate temperature readings for wearables.
CHAPTER FIVE

5.0 CONCLUSION AND RECOMMENDATION


It can be concluded that the sole aim of carrying out the design, analysis
and implementation of a dual sensor heat control system was achieved, in
that the aim was to develop a cheap, affordable, reliable and efficient
temperature monitoring device, which was successfully realized at the end
of the design process. One factor that accounts for the cheapness of the
product was the proper choice of components used. The ones that were
readily available were used, which a close substitute was found for those
that were not readily available. The system involves the design,
construction and testing of a heat monitoring device. It involves the
triggering 'ON' of an alarm whenever the temperature of either of the
devices being monitored exceeds a preset level. A temperature sensor (LM
335) was used to monitor variations in temperature level of the two devices
to which it is connected. At the end of the construction, the work was
tested to find out if the objectives of the research were met and it was
discovered that the research functioned properly according to design. In
the future, we intend to add a digital temperature display into the device to
show the involved temperatures, use a wider range Temperature sensor to
have more industrial importance, and interface the system with a computer
to provide better control and monitoring.
REFERENCES

Anliker, U., Ward, J. a, Lukowicz, P., Tröster, G., Dolveck, F., Baer, M.,
Keita, F., Schenker, E.B., Catarsi, F., Coluccini, L., Belardinelli, A.,
Shklarski, D., Alon, M., Hirt, E., Schmid, R., Vuskovic, M.: AMON: A
Wearable Multiparameter Medical Monitoring and Alert System. IEEE
Trans. Inf. Technol. Biomed. 8(4), 415–27 (2004).
Bakker, A. (2002): CMOS smart temperature sensors—An Overview.
Proceedings of IEEE Sensors 2002. pp. 1423–1427. IEEE (2002).

Bedford, R.E., Bonnier, G., Maas, H., Pavese, F.: Recommended values of
temperature on the International Temperature Scale of 1990 for a selected
set of secondary reference points. Metrologia. 33, 133–154 (1996).
Berger, L.I.: Properties of Semiconductors. In: Haynes, W.M. (ed.) CRC
Handbook of Chemistry and Physics, 94th Edition. pp. 12–80 – 12–93.
CRC Press (2013).
Bergman, T.L., Lavin, A.S., Incropera, F.P., Dewitt, D.P.: Fundamentals of
Heat and Mass Transfer, Seventh Edition. John Wiley & Sons (2011).
Biró, T.S.: How to Measure Temperature. In: Biró, T.S. (ed.) Is There a
Temperature? Conceptual Challenges at High Energy, Acceleration and
Complexity. Fundamentals Theories of Physics, vol 171,. pp. 5–27.
Springer New York, New York, NY (2011).
Dames, C.: Resistance Temperature Detectors. In: Li, D. (ed.)
Encyclopedia of Microfluidics and Nanofludics. pp. 1782–1790.
Springer US (2008).
Desmarais, R., Breuer, J.: How to Select and Use the Right Temperature
Sensor. Sensors Online,
http://archives.sensorsmag.com/articles/0101/24/index.htm, (2001).
Fujinami, K., Xue, Y., Murata, S., Hosokawa, S.: A Human-Probe System
That Considers On-body Position of a Mobile Phone. In: Streitz, N. and
Stephanidis, C. (eds.) Distributed, Ambient, and Pervasive Interactions
- First International Conference, DAPI /HCII 2013. Lecture Notes in
Computer Science, vol. 8028. pp. 99–108. Springer Berlin Heidelberg
(2013).
Habashi, F.: Silicon, Physical and Chemical Properties. In: Kretsinger,
R.H., Uversky, V.N., and Permyakov, E.. (eds.) Encyclopedia of
Metalloproteins. pp. 1998–2000. Springer New York (2013).
Hall, L.A.: Survey of Electrical Resistivity Measurements on 16 Pure
Metals in the Temperature Range 0 to 273K. NBS Tech. Note 365.
February, 1–111 (1968).
Ho, C.Y., Powell, R.W., Liley, P.E.: Heat Conductivity of the Elements. J.
Phys. Chem. Ref.
Data. 1(2), 279–421 (1972).
Janata, J.: Heat Sensors. Principles of Chemical Sensors. pp. 51–62.
Springer US (2009). Johnstone, J. a, Ford, P. a, Hughes, G., Watson, T.,
Garrett, A.T.: Bioharness(TM) Multivariable
Monitoring Device. Part. I: Validity. J. Sports Sci. Med. 11(3), 400–8
(2012).
Kenny, T.: Sensor Fundamentals. In: Wilson, J.S. (ed.) Sensor Technology
Handbook. pp. 1–20.
Elsevier Science & Technology (2004).
Lane, N.D., Miluzzo, E., Lu, H., Peebles, D., Choudhury, T.: A Survey of
Mobile Phone Sensing.
IEEE Commun. Mag. 48(9), 140–150 (2010).
Love, J.: Temperature Measurement. In: Love, J. (ed.) Process Automation
Handbook - A Guide to Theory and Practice. pp. 99–106. Springer US
(2007).
Mekid, S., Starr, A., Pietruszkiewicz, R.: Intelligent Wireless Sensors. In:
Holmberg, K., Adgar, A., Arnaiz, A., Jantunen, E., Mascolo, J., and
Mekid, S. (eds.) E-maintenance. pp. 83–123. Springer London (2010).
Mukhopadhyay, S.C.: Wireless Sensors and Sensors Network. In:
Mukhopadhyay, S.C. (ed.) Intelligent Sensing, Instrumentation and
Measurement. Smart Sensor, Measurement, and Instrumentation, vol. 5.
pp. 55–69. Springer Berlin Heidelberg (2013).
Nicholas, J. V., White, D.R.: Traceable Temperatures - An Introduction to
Temperature Measurement and Calibration, Second Edition. John
Wiley & Sons (2001).
Peterson, K.E.: Silicon as a Mechanical Material. Microelectron. Reliab. 23,
403 (1983).
Preston-Thomas, H.: The International Temperature Scale of 1990 (ITS-
90). Metrologia. 27(3), 3–10 (1990).
Roozeboom, C.L., Hopcroft, M.A., Smith, W.S., Sim, J.Y., Member, S.,
Wickeraad, D.A., Hartwell, P.G., Pruitt, B.L.: Integrated Multifunctional
Environmental Sensors. J. Microelectromechanical Syst. 22(3), 779–793
(2013).
Rotem, E., Hermerding, J., Aviad, C., Harel, C.: Temperature
Measurement in the Intel Core Duo Processor. Procceedings of the
12th International Workshop on Heat Investigations, Therminic, Nice
2006. EDA Publishing Association (2006).
Serway, R.A.: Physics: For Scientists and Engineers with Modern Physics,
3rd Edition. Saunders College Publishing, Philedelphia (1990).

Sharifi, S., Liu, C., Rosing, T.S.: Accurate Temperature Estimation for
Efficient Heat Management. 9th Int. Symp. Qual. Electron. Des. 137–
142 (2008).
Talavera, G., Martin, R., Rodríguez-alsina, A., Garcia, J., Fernández, F.,
Carrabina, J.: Protecting Firefighters with Wearable Devices. In: Bravo,
J., López-de-Ipiña, D., and Moya, F. (eds.) Ubiquitous Computing and
Ambient Intelligence. Lecture Notes in Computer Science, vol. 7657.
pp. 470–477. Springer Berlin Heidelberg (2012).
Tomsen, V.: Response Time of a Thermometer. Phys. Teach. 36, 540–541
(1998).
Touloukian, Y.S., Kirby, R.K., Taylor, R.E., Desai, P.D.: Heat Expansion -
Metallic Elements and Alloys. In: Touloukian, Y.S., Kirby, R.K.,
Taylor, R.E., and Desai, P.D. (eds.) The TPRC Data Series, vol 12.
Plenum Publishing Corporation (1975).
Wagman, D.D., Evans, W.H., Parker, V.B., Schumm, R.H., Halo, I.,
Bailey, S.M., Churney, K.L., Nuttall, R.L.: The NBS Tables of
Chemical Thermodynamic Properties. J. Phys. Chem. Data. 11, Supp.,
1–392 (1982).
Yacobi, B.G.: Semiconductor Materials : An Introduction to Basic
Principles. Kluwer Academic Publishers, Secaucus, NJ (2002).
The RTD. Omega Engineering Inc.,
http://www.omega.com/temperature/Z/TheRTD.html (2014) Accessed
4 May 2014.
ANSI and IEC Color Codes for Thermocouples, Wire and
Connectors/Thermocouple Tolerances. Omega Engineering Inc.,
http://www.omega.com/temperature/pdf/tc_colorcodes.pdf (2014)
Accessed 05 May 2014.
Electrical Conductivity and Resistivity. NTD Resource Center.,
http://www.ndt-ed.org/Education
Resources/CommunityCollege/Materials/Physical_Chemical/Electrical.
htm (2014) Accessed
04 June 2014.
Physical Properties of Thermoelement Material. Omega Engineering Inc.,
http://www.omega.com/ temperature/Z/pdf/z016.pdf (2014) Accessed 04
May 2014.

You might also like