Download as doc, pdf, or txt
Download as doc, pdf, or txt
You are on page 1of 44

THERMOELECTRIC POWER GENERATION - 12 MOST FREQUENTLY

ASKED QUESTIONS

Dowload the 43 Most Frequently Asked Questions about


Thermoelectric Power Generation

1. How does thermoelectric technology work?

This gets complicated—but then if it was easy, everybody


would be explaining it. We’ll just take it one small step at a
time.

First, we need to understand the relationship between the flow


of electricity and the existence of charge carriers. According to
electron theory, electricity is the movement of electrons in a
circuit. It occurs whenever there is a continuous conductive
path across an applied voltage. The voltage provides an
electromotive force which sets the electrons into motion. The
resulting electrical current is measured in terms of the number
of electrons moving past a given point in one second, where
one ampere (or ‘amp’ for short) equals the movement of 6.25 x
1018 ( ten to the 18th power) electrons per second.

figure 1

Charge carriers are the physical components of a material


which allow it to conduct electricity. The precise nature of
these carriers, is a function of the material’s atomic structure.
In the simplest examples, like copper, the material is a pure
element which has only a single valence electron in its outer
shell (see Figure 1). The fewer the number of electrons in an
element’s outer shell, the more loosely bound it is to the
atom’s nucleus, and the easier it is to make it flow with the
application of a voltage. Other elements with a single outer
electron include silver and gold and they are excellent
conductors.

figure 2

Conductivity takes a somewhat different form when it comes to


semiconductor material. For electronic applications,
semiconductor materials are ‘grown’ into crystalline structures
which are given conductive properties by virtue of the
impurities (or dopants) which are added. In their purest form
(i.e., without dopants), the base semiconductor materials form
crystalline lattices which become very stable by sharing
electrons among the constituent atoms. Figure 2 shows such a
configuration for a silicon crystal. In looking at the shell
mapping, be aware that the electrons (shown in red), are
actually in constant motion as they orbit the nuclei in the
lattice. The shared electrons, however, are continually pulled
into the orbits of adjacent nuclei to maintain the structural
stability of the lattice. In this pure state, the material is not
very conductive.

figure 3

Once the impurities are added to the mix, however, the


conductive properties are radically affected. For example, if we
have a crystal formed primarily of silicon (which has four
valence electrons), but with arsenic impurities (having five
valence electrons) added, we wind up with ‘free’ electrons
which do not fit into the crystalline structure (see Figure 3).
These electrons are thus ‘loosely bound’ and when a voltage is
applied, they can be easily set in motion to allow electrical
current to pass. The loosely bound electrons are considered
the charge carriers in this ‘negatively doped’ material (which
is referred to as ‘N’ material).

figure 4

It is also possible to form a more conductive crystal by adding


impurities which have one less valence electron. For example,
if Indium impurities (which have three valence electrons) are
used in combination with silicon, this creates a crystalline
structure which has ‘holes’ in it—that is, places within the
crystal where an electron would normally be found if the
material was pure. These ‘holes’ make it much easier to
convey electrons through the material upon the application of
a voltage. In this case, ‘holes’ are considered to be the charge
carriers in this ‘positively doped’ conductor (which is referred
to as ‘P’ material).

It is critical here to understand that the existence of charge


carriers is entirely a property of a given material. The vast,
vast majority of conductors—including those employed to make
electrical connections—use electrons as the charge carriers
and would be considered ‘N’ material. ‘P’ material can only be
fabricated within crystalline structures.

Okay, now that we have a basic understanding of electricity


and the nature of charge carriers, we need to come to grips
with an important concept in power generation. Sometimes it
is possible to set charge carriers in motion through interaction
with other energy sources. For instance, if a magnetic field is
moved along a conductor, the effect of that field upon the
electrons (assuming that there is a complete path), will cause
electrical current to flow. In essence, if you can force charge
carriers to move, you can create voltage and current flow. This
is not only true when there is an interaction between charge
carriers and magnetic fields, but when those carriers are set in
motion by the flow of heat.
Thus we come to the nitty gritty of Seebeck technology.
Whenever an electrical conductor is strung between two
different temperatures, the conductor is capable of
transferring thermal energy from the warmer side to the
colder one. Furthermore, the physical process of transferring
that heat, also tends to move electrical charge carriers within
the conductor in the same direction as the heat. Conceivably
then, this charge carrier movement can be used to generate
electrical current—if we can find a way to effectively complete
the circuit.

Here, however, we run up against a major issue. If the


conductor which completes the circuit is identical to the first
conductor, the flow of thermal energy will create a potential
for equal charge carrier movement in both conductors.
Furthermore, the potential for current flow in one conductor is
in complete opposition to that in the other conductor. The
result is no net current flow.

If we employ two dissimilar conductors, on the other hand, we


get quite a different result. With differing capacities for
moving charge carriers in response to thermal flow, the
current level in one conductor will overcome (or in some cases,
complement) the potential for thermally-generated current
flow in the other conductor. The net effect is a continuous
current level which is equal to the generated current capacity
of the primary conductor (for the given temperature
difference) minus the generated current capacity of the second
conductor. The existence of this net current flow, indicates that
a voltage is created through the movement of heat and we can
get a direct measurement of this voltage level by breaking the
circuit and measuring across the opened terminals with a
voltmeter. Note that the ability of two dissimilar conductors to
produce a voltage when a temperature difference is applied, is
called the Seebeck effect. The voltage which results is referred
to as Seebeck voltage.

Probably the most well-known example of this phenomenon, is


the common thermocouple. For example, with a K-type
thermocouple made of two wires—one composed of a nickel-
chromium alloy and the other from nickel-aluminum, if one
junction is at 100° C and the other junction (the so-called
‘reference junction’) is at 0° C, a voltage of approximately
4.096 millivolts is produced. In general, the voltage generated
by a thermocouple is a function of two things: 1) the
temperature difference (DT) between the two thermocouple
junctions, and 2) the nature of the conductors employed
(including their temperature dependencies).

Of course, thermocouples are used primarily for temperature


measurement—not power generation. Thermoelectric power
generation (TEG) devices typically use special semiconductor
materials which are optimized for the Seebeck effect. The
circuit shown in Figure 10 demonstrates the simplest possible
example. It shows a single ‘N’-type semiconductor pellet
connected across a voltmeter.

figure 10

figure 11

As the heat moves from the hot to the cold side of the pellet,
the charge carriers (i.e., electrons from the dopants) are
carried with the heat. Heat also effects charge carrier
movement in the return path (typically copper wire). Because
the heat movement can carry far more charge carriers in the
semiconductor material than in the circuit’s return path,
however, a significant potential difference (i.e., Seebeck
voltage) is generated. In this example, the Seebeck voltage
would be about 20 mV.

In thermoelectric power generation, ‘P’ pellets are also


employed. Figure 11 shows a basic configuration. Note how the
flow of electrons goes in a direction opposite to that of hole
flow. It is through the use of both N and P type materials in a
single power generation device, that we can truly optimize the
Seebeck effect. As shown in Figure 12, the N and P pellets are
configured thermally in parallel, but electrically in a series
circuit. Because electrical current (i.e., moving electrons) flows
in a direction opposite to that of hole flow, the current
generating potentials in the pellets do not oppose one another,
but are series-aiding. Thus, if each pellet developed a Seebeck
voltage of 20 mV, this combination of an N and P pellet would
generate approximately 40 mV rather than zero volts.

figure 12

Of course, in truly practical Thermoelectric power generation,


many such P & N couples are employed to bring the Seebeck
voltage up to useful levels. The illustration in Figure 13 shows
a three-couple device (more typically, a Seebeck module would
have 127 couples or more). Note the direction of electrical
current flow in the N/P series configuration (assuming a load is
connected across the Seebeck device).

figure 13

2. Do thermoelectric power generation products employ silicon-based


semiconductor material?

They can. Tellurex, however, uses bismuth/telluride structures


to optimize performance. While similar dopants are employed
in both semiconductor technologies, the crystalline latices
which form from Bismuth/Telluride, are far more complex. The
same principles of ‘N’ and ‘P’ material apply, though.
3. How is a typical thermoelectric power generation system
configured?

Fundamentally, there are four basic components: a heat


source, a thermoelectric power generation module (i.e., a
thermoelectric generator—also known as a Seebeck device), a
‘cold-side’ heat sink, and the electrical load. The system may
also include a voltage regulation circuit, or a fan for the heat
sink. The illustration in Figure 14 shows one example.
In this case we have a burner box with a propane fuel source.
It is shown with the burner box open on one end, but in reality,
it would be enclosed. The thermoelectric module is then
sandwiched between the heat source and the cold-side sink.
While this example shows only a single thermoelectric power
generation module, in reality, several modules might be
deployed in whatever series/parallel electrical arrangement
best served the load.

figure 14
4. Do I have to use a heat sink in my design?

It would be virtually impossible to get an adequate DT without


some type of heat sink. However, you can sometimes reduce
the size requirement for the sink (i.e., fin surface area) if you
can find a way to insure good air flow.

5. Are any special precautions required for the hot side of the
system?

Yes. First and foremost, you want to prevent the hot-side


temperature of the thermoelectric device from exceeding the
melting temperature of the solder employed to secure the
semiconductor pellets to the copper tabs. It is recommended
that the temperature be kept below 200° C. Toward this end, it
is a good idea to use some type of ‘heat spreader’ to prevent
hot spots at the hot-side module interface. Usually this means
employing a relatively thick casting or extrusion between the
heat source and the module.

On the mechanical side—especially when using multiple


devices—you need to find a means of applying compression
between the hot and cold sides, which will apply even pressure
across the thermoelectric modules and, most importantly,
prevent the hot-side interface from bowing. If there is too
great an expanse between compression points, the hot side
interface can distort to the point where some modules are
crushed or the thermal interface is compromise.

6. What does the specification, THot, mean?

This is the temperature at the mounting surface of the


thermoelectric module, which comes in contact with the heat
source (i.e., the hot side of the system).

7. What does the specification, TCold, mean?

This is the temperature at the mounting surface of the module,


which comes in contact with the cold-side heat sink.

8. What does ‘no-load voltage’ (VNL) mean?

This is the voltage output of the thermoelectric power


generation system when no electrical load is connected.

9. What does ‘load voltage’ (VL) mean?


This is the voltage output of the TEG system when an electrical
load is connected.

10. What does internal resistance (RInt) mean?

This is the electrical resistance of the thermoelectric power


generation module (or module array).

11. What does ‘power conversion efficiency’ mean?

It is the ratio of power output to power input, expressed as a


percentage. In this case, power output would be the wattage
dissipated in the electrical load and power input would be the
rate of energy use (e.g., watts, BTU’s/hr) to create the
necessary DT.
12. What does ‘electrical efficiency’ mean?

It is the ratio of electrical power dissipated in the load to the


total amount of power generated (including the dissipation in
the internal resistance).

WORKING OF PELTIER PLATE

WORKING MODEL

PELTIER PLATE
HOT SIDE COLD SIDE

COPPER PLATES

Thermoelectric Power Generation Technology FAQ


©2006, Tellurex Corporation
1462 International Drive
Traverse City, Michigan 49684
231-947-0110 • www.tellurex.com
The Most Frequently Asked Questions About
Thermoelectric Power Generation Technology
© 2006 Tellurex Corporation
1. How does this technology work?
This gets complicated—but then if it was easy, everybody would be explaining it.
We’ll just take it one small
step at a time.
First, we need to understand the relationship between the flow of electricity and the
existence of charge carriers.
According to electron theory, electricity is the movement of electrons in a circuit. It
occurs whenever there is
a continuous conductive path across an applied voltage. The voltage provides an
electromotive force which
sets the electrons into motion. The resulting electrical current is measured in terms
of the number of electrons
moving past a given point in one second, where one ampere (or ‘amp’ for short)
equals the movement of
6.25 x 1018 electrons per second.
Charge carriers are the physical components of a
material which allow it to conduct electricity. The
precise nature of these carriers, is a function of the
material’s atomic structure. In the simplest examples,
like copper, the material is a pure element which has
only a single valence electron in its outer shell (see
Figure 1). The fewer the number of electrons in an
element’s outer shell, the more loosely bound it is to
the atom’s nucleus, and the easier it is to make it flow
with the application of a voltage. Other elements with
a single outer electron include silver and gold and
they are excellent conductors.
Conductivity takes a somewhat different form when
it comes to semiconductor material. For electronic
applications, semiconductor materials are ‘grown’
into crystalline structures which are given conductive
properties by virtue of the impurities (or dopants)
which are added. In their purest form (i.e., without
dopants), the base semiconductor materials form
crystalline lattices which become very stable by
sharing electrons among the constituent atoms.
Figure 2 shows such a configuration for a silicon
crystal. In looking at the shell mapping, be aware
that the electrons (shown in green), are actually in
constant motion as they orbit the nuclei in the lattice.
The shared electrons, however, are continually pulled
into the orbits of adjacent nuclei to maintain the
structural stability of the lattice. In this pure state,
the material is not very conductive.
Once the impurities are added to the mix, however,
the conductive properties are radically affected. For
example, if we have a crystal formed primarily of
silicon (which has four valence electrons), but with
arsenic impurities (having five valence electrons)
added, we wind up with “free” electrons which do not
fit into the crystalline structure (see Figure 3). These
Single valence electron
(outer shell)
Electron Shells
(2,8,18,1)
Atom’s Nuceus
(neutrons & protons)
Figure 1: Structure of copper atom.
Only outer shells shown
All electrons in outer
shell are shared among
atoms
Figure 2: With four electrons in the outer shell, silicon atoms form a
stable cyrstalline lattice by sharing electrons with each other.
Arsenic atom
Silicon atom
Free electron from
arsenic atom
Figure 3: Arsenic dopant adds free electrons to the crystalline lattice,
making it more electrically conductive, creating ‘N’ material.
Thermoelectric Power Generation Technology FAQ
electrons are thus “loosely bound” and when a voltage is applied, they can be easily
set in motion to allow
electrical current to pass. The loosely bound electrons are considered the charge
carriers in this ‘negatively
doped’ material (which is referred to as ‘N’ material).
It is also possible to form a more conductive
crystal by adding impurities which have one less
valence electron. For example, if Indium impurities
(which have three valence electrons) are used in
combination with silicon, this creates a crystalline
structure which has “holes” in it—that is, places
within the crystal where an electron would normally
be found if the material was pure. These holes
make it much easier to convey electrons through the
material upon the application of a voltage. In this
case, holes are considered to be the charge carriers
in this “positively doped” conductor (which is
referred to as ‘P’ material).
It is critical here to understand that the existence of charge carriers is entirely a
property of a given material.
The vast, vast majority of conductors—including those employed to make electrical
connections—use
electrons as the charge carriers and would be considered ‘N’ material. ‘P’ material
can only be fabricated within
crystalline structures.
Okay, now that we have a basic understanding of electricity and the nature of charge
carriers, we need to come
to grips with an important concept in power generation. Sometimes it is possible to
set charge carriers in motion
through interaction with other energy sources. For instance, if a magnetic field is
moved along a conductor, the
effect of that field upon the electrons (assuming that there is a complete path), will
cause electrical current to
flow. In essence, if you can force charge carriers to move, you can create voltage and
current flow. This is not
only true when there is an interaction between charge carriers and magnetic fields,
but when those carriers are
set in motion by the flow of heat.
Thus we come to the nitty gritty of Seebeck technology. Whenever an electrical
conductor is strung between
two different temperatures, the conductor is capable of transferring thermal energy
from the warmer side
to the colder one. Furthermore, the physical process of transferring that heat, also
tends to move electrical
charge carriers within the conductor
in the same direction as the heat (see
Figure 5). Conceivably then, this
charge carrier movement can be used to
generate electrical current—if we can
find a way to effectively complete the
circuit.
Here, however, we run up against a
major issue. If the conductor which
completes the circuit is identical
to the first conductor, the flow of
thermal energy will create a potential
for equal charge carrier movement
in both conductors (see Figure 6).
Furthermore, the potential for current
flow in one conductor is in complete
opposition to that in the other
conductor. The result is no net current
flow.
If we employ two dissimilar
conductors, on the other hand, we get
quite a different result (see Figure
Indium atom
‘Hole’ in latice
Silicon atom
Figure 4: Indium dopant leaves “holes” in the crystalline lattice,
making it more electrically conductive while creating ‘P’ material.
Warmer side Colder side
Potential charge carrier movement
Conductor
Thermal energy movement
Figure 5: The capacity for charge carrier movement is created
whenever thermal energy moves through an electrical conductor.
Figure 5: The capacity for charge carrier movement is created whenever thermal
energy moves through an electrical conductor.
Warmer junction
Colder junction
Potential charge carrier movement
Potential charge carrier movement
Conductor #1
Conductor #2
Figure 6: When identical charge carriers are employed, there is a complete circuit,
but the two potential differences cancel one another (resulting in no net current flow). Figure 6: When identical
charge carriers are employed, there is a complete circuit,
but the two potential differences cancel one another (resulting in no net current flow).
Thermoelectric Power Generation Technology FAQ
7). With differing capacities for
moving charge carriers in response to
thermal flow, the current level in one
conductor will overcome (or in some
cases, complement) the potential for
thermally-generated current flow in
the other conductor. The net effect
is a continuous current level which is
equal to the generated current capacity
of the primary conductor (for the
given temperature difference) minus
the generated current capacity of the
second conductor. The existence
of this net current flow, indicates
that a voltage is created through the
movement of heat and we can get a
direct measurement of this voltage
level by breaking the circuit and
measuring across the opened terminals
with a voltmeter (see Fig. 8). Note
that the ability of two dissimilar
conductors to produce a voltage when
a temperature difference is applied, is
called the Seebeck effect. The voltage
which results is referred to as Seebeck
voltage.
Probably the most well-known example
of this phenomenon, is the common
thermocouple (see Figure 9). For
example, with a K-type thermocouple
made of two wires—one composed of
a nickel-chromium alloy and the other
from nickel-aluminum, if one junction
is at 100° C and the other junction (the
so-called ‘“reference junction”) is at
0° C, a voltage of approximately 4.096
millivolts is produced. In general, the
voltage generated by a thermocouple
is a function of two things: 1) the
temperature difference (⊗T) between
the two thermocouple junctions, and 2) the nature of
the conductors employed (including their temperature
dependencies).
Of course, thermocouples are used primarily for
temperature measurement—not power generation.
Thermoelectric power generation (TEG) devices
typically use special semiconductor materials which
are optimized for the Seebeck effect. The circuit
shown in Figure 10 demonstrates the simplest possible
example. It shows a single ‘N’-type semiconductor
pellet connected across a voltmeter. As the heat moves
from the hot to the cold side of the pellet, the charge
carriers (i.e., electrons from the dopants) are carried with the heat. Heat also effects
charge carrier movement
in the return path (typically copper wire). Because the heat movement can carry far
more charge carriers in the
Warmer
junction
Colder
junction
Figure 8: Breaking the circuit, allows for the measurement
of the Seebeck voltage which is produced.

V
Warmer junction
Colder junction
Figure 7: Use of two dissimilar conductors creates a complete circuit which
allows for continuous current flow. The conductor with the greatest capacity
for charge carrier movement will determine the direction of current flow.
Figure 7: Use of two dissimilar conductors creates a complete circuit which allows
for continuous current flow. The conductor with the greatest capacity for charge
carrier movement will determine the direction of current flow.

V N
flo
heat
flow
Hot Side
Cool Side
Charge
Carriers
(electrons)
Electron Flow
Electron Flow
-
+
Figure 8: Breaking the circuit, allows for the measurement of the Seebeck voltage
which is produced.
Figure 9: Thermocouple using 0° C reference junction.
100° C Measurement
junction
0° C reference
junction (ice bath)
Figure 9: Thermocouple using 0° C reference junction.

V
4.096 mV
Figure 10
Thermoelectric Power Generation Technology FAQ
semiconductor material than in the circuit’s return path,
however, a significant potential difference (i.e., Seebeck
voltage) is generated. In this example, the Seebeck
voltage would be about 20 mV.
In thermoelectric power generation, ‘P’ pellets are also
employed. Figure 11 shows a basic configuration. Note
how the flow of electrons goes in a direction opposite to
that of hole flow.
It is through the use of both N and P type materials
in a single power generation device, that we can truly
optimize the Seebeck effect. As shown in
Figure 12, the N and P pellets are configured
thermally in parallel, but electrically in a series
circuit. Because electrical current (i.e., moving
electrons) flows in a direction opposite to that
of hole flow, the current generating potentials
in the pellets do not oppose one another, but are
series-aiding. Thus, if each pellet developed a
Seebeck voltage of 20 mV, this combination of
an N and P pellet would generate approximately
40 mV rather than zero volts.
Of course, in truly practical TEG’s, many
such P & N couples are employed to bring the
Seebeck voltage up to useful levels. The illustration in Figure 13 shows a three-
couple device (more typically, a
Seebeck module would have 127 couples or more). Note the direction of electrical
current flow in the N/P series
configuration (assuming a load is connected across the Seebeck device).
2. Do TEG’s employ silicon-based semiconductor material?
They can. Tellurex, however, uses bismuth/telluride structures to optimize
performance. While similar dopants
are employed in both semiconductor technologies, the crystalline latices which form
from Bismuth/Telluride,
are far more complex. The same principles of ‘N’ and ‘P’ material apply, though.

V P
flo
heat
flow
Hot Side
Cool Side
Charge
Carriers
(holes)
Electron Flow
Electron Flow
-
+
Figure 11

NP
Hot Side Hot Side
Cool Side
Charge Carrier
Movement
(electrons)
Charge Carrier
Movement
(holes)
+-
Thermal Energy Movement
Figure 12
Thermal
Flow
Hot Side
Cool Side
+-
NPNPNP
Electron Current Electron Current
Electron Current Electron Current
Figure 13
Thermoelectric Power Generation Technology FAQ
3. How is a typical TEG system
configured?
Fundamentally, there are four basic components: a
heat source, a TEG module (i.e., a thermoelectric
generator—also known as a Seebeck device), a
“cold-side” heat sink, and the electrical load. The
system may also include a voltage regulation
circuit, or a fan for the heat sink. The illustration
in Figure 14 shows one example.
In this case we have a burner box with a propane
fuel source. It is shown with the burner box open
on one end, but in reality, it would be enclosed.
The TE module is then sandwiched between the
heat source and the cold-side sink. While this
example shows only a single TEG module, in
reality, several modules might be deployed in
whatever series/parallel electrical arrangement best served the load.
4. Do I have to use a heat sink in my design?
It would be virtually impossible to get an adequate ⊗T without some type of heat
sink. However, you can sometimes
reduce the size requirement for the sink (i.e., fin surface area) if you can find a way
to insure good air flow.
5. Are any special precautions required for the hot side of the system?
Yes. First and foremost, you want to prevent the hot-side temperature of the TE
device from exceeding the
melting temperature of the solder employed to secure the semiconductor pellets to
the copper tabs. It is
recommended that the temperature be kept below 200° C. Toward this end, it is a
good idea to use some type of
“heat spreader” to prevent hot spots at the hot-side module interface. Usually this
means employing a relatively
thick casting or extrusion between the heat source and the module.
On the mechanical side—especially when using multiple devices—you need to find a
means of applying
compression between the hot and cold sides, which will apply even pressure across
the modules and, most
importantly, prevent the hot-side interface from bowing. If there is too great an
expanse between compression
points, the hot side interface can distort to the point where some modules are
crushed or the thermal interface is
compromised.
6. What does the specification THot mean?
This is the temperature at the mounting surface of the module, which comes in
contact with the heat source (i.e.,
the hot side of the system).
7. What does the specification TCOLD mean?
This is the temperature at the mounting surface of the module, which comes in
contact with the cold-side heat
sink.
8. What does “no-load voltage” (VNL) mean?
This is the voltage output of the TEG system when no electrical load is connected.
9. What does “load voltage” (VL) mean?
This is the voltage output of the TEG system when an electrical load is connected.
10. What does “internal resistance” (RInt) mean?
This is the electrical resistance of the TEG module (or module array).
Fuel
Burner Box
Exhaust
Electrical Output
Cold Side
Heat Sink
TE Module
Figure 14
Thermoelectric Power Generation Technology FAQ
11. What does “power conversion efficiency” mean?
It is the ratio of power output to power input, expressed as a percentage. In this case,
power output would be the
wattage dissipated in the electrical load and power input would be the rate of energy
use (e.g., watts, BTU’s/hr)
to create the necessary ⊗T.
12. What does “electrical efficiency” mean?
It is the ratio of electrical power dissipated in the load to the total amount of power
generated (including the
dissipation in the internal resistance).
13. What does “worst case operating point” mean?
Within the range of possible operating circumstances, the worst case will occur at the
point where the generating
system can just meet the expected demands of the electrical load. In most systems,
this will be when: 1) the
generator is at the lowest expected ⊗T, and 2) the load requires the greatest
expected current draw. Some
systems may have a worst case operating point under other unique conditions,
however. The key concept is that
the load requires everything that the generator can muster.
14. What sort of efficiencies typically result from TE power generation?
To put it bluntly, TEG’s are not used for their incredible power conversion efficiency.
When your primary
design goal is maximizing efficiency, you’re plainly going to choose another
technology. Generally, if you’re
getting 5 to 10% between power in and power out on a TEG, you’re doing pretty well.
As for the efficiency of
any specific TEG system . . . well, that depends. The properties of thermoelectric
modules are very temperature
dependent and efficiencies can vary widely depending on the operating parameters
of the system at hand.
Generally speaking, the higher the ⊗T and the closer you get to a point of maximum
power transfer (matching
the generator to your load—more on this to come), the more efficient your system
will be with respect to power
conversion.
15. Why, then, would I want to use a thermoelectric generator (TEG) as opposed
to some other approach?
Thermoelectric power generation is definitely a niche technology. TEG use is most
viable in applications
where waste heat can be converted into usable electric power. In these situations,
the input power is essentially
free—heat that is a by-product of some other necessary process. As long as the
opposite side of the module can
be effectively cooled (usually with a passive heat sink) to create a ⊗T across the TEG,
there is the potential for
extracting DC to power an electrical load.
Another common justification for using TEG’s, is when there is an overriding reason
precluding the use of
other technologies. For example, the remoteness of a use site may make it
impractical for someone to maintain
a gas-powered mechanical generator. As long as a source of fuel is readily available,
a properly-operated
thermoelectric generator can run for a very long period without human attention.
16. What is “maximum power transfer” and
why is it so important in designing TEG
systems?
It’s too bad there’s not a short answer to this question, but
alas. People are so accustomed to working with regulated
power sources these days, that they hardly pay attention
to this very fundamental concept of electronics. With
a regulated power supply, you can connect all kinds of
different loads (within some limits), and the voltage will
stay relatively constant. You can even buy constant current
sources that will keep the current steady despite variations
in load.
TEG’s just don’t work like regulated power supplies. At
any given ⊗T, as the load resistance decreases, so does the
output voltage; as the load resistance increases, the output
Load
INT
R
R
No-load
voltage
TEG
Model
Figure 15
Thermoelectric Power Generation Technology FAQ
voltage will follow suit. There is no inherent output regulation (although it can be
provided with additional
circuitry). Why? Because TEG’s have appreciable “internal resistance”. This means
that as more current is
drawn by the electrical load (e.g., as load resistance decreases), more of the
available power is dissipated within
the TEG; furthermore there is an appreciable voltage drop across this internal
resistance.
To understand the electrical model for a TEG system, imagine the no-load voltage
(i.e., the open-circuit voltage
output of a TEG) applied to a series circuit consisting of the module’s internal
resistance (RINT) and the electrical
load (RLoad). As in any series circuit, the voltages will ‘drop’ in proportion to the
resistances. Thus, if the noload
voltage was 3 VDC, the internal resistance was 3 ∧, and the load was 6 ∧, there would
be a 2 VDC drop
across the load.
So how does “maximum power transfer” fit into all of this? Here’s the deal. For any
electrical circuit, you will
transfer the most power to the load when the resistance of the load equals the
internal resistance of the voltage
source. Note we are not saying that this point of operation will give you the
maximum amount of voltage or
current—that is simply not true—but you will derive the greatest power output at that
point. For example, in the
circuit with the no-load voltage of 3 VDC and 3 Ω of internal resistance, we could use
the following formulas to
determine load voltage, current, and power for various load resistances:
Table 1 shows what would result for a range of load resistance values
between 0.5 and 8 Ω. While the load voltage increases with load resistance
and load current increases as the load resistance decreases, it can be seen
that the maximum amount of power is transferred to the load when the load
resistance equals the internal resistance of the TEG. This is an immutable
principle which can actually be proven as the general case using advanced
mathematics. It can also be demonstrated empirically in the lab. Ironically,
at this operating point, you will dissipate equal amounts of power within
both the load and the internal resistance of the generator. This, no doubt,
seems less than ideal, but if you want to get the maximum power output
from your generator, you can only do so with a matched load.
Okay, so we just match our generator and load, right? Matching loads,
however, can get fairly tricky. It is most easily done when sales quantities
can justify a custom TEG module; then the pellet configuration can be
optimized for the specific application. When employing off-the-shelf
options, on the other hand, the designer is left to arrive at a series/parallel
module combination that can do the best job of load-matching with the
fewest devices. Further complicating matters in many cases, will be the
fact that operating conditions may change over time and your design may
be matched for one set of circumstances but not for another. If the ambient
environment is dynamic and/or the load resistance is variable, therefore, the
engineer must design to the worst case and then manage surplus capacity with a
regulatory circuit.
17. Why is it so important to approximate maximum power transfer in a TEG
system design?
This is where we get back to efficiency—both in terms of power conversion and
economy of design. Let’s look
at the latter criteria first.
RLoad VL IL PL
0.5 0.43 0.86 0.367
1 0.75 0.75 0.563
1.5 1.00 0.67 0.667
2 1.20 0.60 0.720
2.5 1.36 0.55 0.74
3 1.50 0.50 0.750
3.5 1.62 0.46 0.746
4 1.71 0.43 0.735
4.5 1.80 0.40 0.720
5 1.88 0.38 0.703
5.5 1.94 0.35 0.685
6 2.00 0.33 0.667
6.5 2.05 0.32 0.648
7 2.10 0.30 0.630
7.5 2.14 0.29 0.612
8 2.18 0.27 0.595

() ( VDC) VDC
RR
VVR
INT Load
Load
Load NL 2
36
63= 


 

∧+∧
∧=  


 

+
=
( ) ( )( ) LLL
Load
L
L
INT Load
Load
L NL PVI
R
IV
RR
RVV= =  


 

+
=
Thermoelectric Power Generation Technology FAQ
In any product, it is important to keep costs down. Toward that end in TEG design,
you must accept the reality
that when the internal resistance of your generator is properly matched to your load,
you will get the job done
with the fewest TEG devices. If you don’t concern yourself with load matching, on the
other hand, it simply
means that you will use more devices and heat sink capacity to generate the
necessary power.
In those systems which are not based upon waste heat, power conversion efficiency
becomes an issue. Here,
given the natural limits of these devices, you want to be as efficient as possible or
your operating expenses (i.e.,
cost of fuel) will be significantly higher. You can only optimize your efficiency to the
extent that you match
your load and get close to achieving maximum power transfer.
Practically speaking, most TEG system designs will be at least somewhat
mismatched. When you’re dealing
with discrete building blocks (i.e., individual TEG modules), you can only approximate
a given resistance with
series/parallel combinations. Generally, it is better to err on the side of having load
resistance exceed internal
resistance (efficiency is a much more slippery slope when it is the other way around),
but there are certainly
exceptions.
18. Is a TEG’s resistance fairly constant over its operating range?
These devices are significantly temperature-dependent and that presents us with
additional challenges in arriving at
design solutions. For example, a module that has an internal resistance of 5Ω at an
average temperature of 100° C,
might reach 6.1Ω at 150° C. This can result in significant shifts in output voltage,
current, and power as operating
temperatures change. For instance, with such a shift in operating temperatures, if the
load was 5 Ω and the no-load
voltage was held at 8 VDC, given the change in module resistance, the load voltage
would decrease from 4 V to 3.6
V, current would drop from 0.8 to 0.72 amps, and power would go from 3.2 to 2.6
watts (nearly a 19% decrease) at
the higher temperatures. Therefore, if there is any potential for variation in operating
conditions, you have to look at
the range of module resistance values that may result and make sure that you can
accommodate them. This usually
means designing to the worst case—making sure that you get sufficient power when
⊗T is low and current demand
is high—and manage the excess generating capacity at all other operating points.
19. What sort of ⊗T is required to generate power?
Any ⊗T at all will result in power generation. The greater the ⊗T, however, the more
power can be derived—all
other things being equal. Put differently, the greater the ⊗T you achieve, the fewer
TEG devices it will take to
power a given load.
20. How big can these devices get?
Theoretically, there is no limit, but practicality does impose some restrictions. Issues
related to thermal
expansion/contraction—and cost—tend to keep module sizes down to a moderate
size. Typical devices range
up to 40 mm (1.57”) square and about 4 mm thick, but there are exceptions. In the
general case, when extra
generating capacity is required, multiple TEG devices will be employed rather than
fabricating some sort of
gargantuan module.
21. How small can these devices get?
Here again, the theoretical limit goes far beyond what is practical. One issue is
manufacturability. At a certain
point, smaller devices become more expensive to make because they are less
adaptable to automated techniques
and require more hand assembly under demanding conditions.
22. Is there any advantage to be gained from
using multi-stage (i.e., layered) TEG’s?
Not really. This sort of series thermal configuration would actually lower the ⊗T
applied to each stage (halving
it in the case of two layers). This, in turn, would reduce the power output of each
stage. There is really nothing
to be gained with this approach.
23. If I need more voltage than can be provided
by a single device, what are my options?
If the electrical current level from a single device is adequate, but the voltage is
insufficient, you usually place
additional modules electrically in series.
Thermoelectric Power Generation Technology FAQ
24. If I need more current than can be provided
by a single device, what are my options?
If the voltage level from a single device is adequate, but the current level is
insufficient, you usually place
additional modules electrically in parallel.
25. What if I need more voltage and current?
Here there is no one clear answer. This requires circuit analysis on various series,
parallel, and series/parallel
combinations of TEG’s until one is found which best matches the load conditions. Of
course, you need to keep
the temperature dependence of the modules in mind as you explore this.
26. How can I determine the optimum series/parallel
configuration for a given load?
Frankly, this is probably best done by presenting your design needs to a Tellurex
sales representative. At
Tellurex, we employ computer models to help us converge on a solution that will
meet your worst-case
requirements and give you extra capacity at other times. To help you with your
application, we will need to
know:
1) The expected range for THot,
2) The lowest expected ⊗T,
3) The desired voltage level,
4) The greatest expected current draw at that voltage level,
5) Any physical limitations which might impinge on the number of Seebeck devices
employed,
6) Whether or not voltage regulation is important (i.e., whether your load can
withstand fluctuations in voltage
and current), and
7) Whether quantities could justify the use of a custom device.
27. What happens if my system is not in a matched-load condition?
Well, it won’t be the end of the world or anything. In fact, most good designs only
come close to matchedload
operation. Furthermore, if the operating conditions (i.e., temperatures, load
resistances, etc.) are variable,
you will be functioning at a mismatch at least some of the time. It should be clear by
now that, whenever your
system is unmatched, your power-conversion efficiency is going to suffer. Unless
efficient use of input power is
critical, the more important objective will be whether or not you are deriving
adequate power for your load.
28. If I need a steady output voltage, do I need some sort of regulator?
Unless your operating conditions are absolutely constant, you
will need a regulator. The instinctive response in this kind of
situation, is to employ a series regulator. Naturally, according
to Murphy’s Law, that is precisely the wrong thing to do. The
problem with series regulators, is that their input voltage must
be at least 1.5 to 2 V higher than the desired output. This
means that you might have to use additional generator capacity
just to deal with the voltage drop across the regulator. Ideally,
you would like a regulator that would allow you to use every
last bit of available voltage (if it is required).
Fortunately, there is a regulator circuit which can deliver on this frontier—the shunt
regulator. A shunt regulator
uses a power semiconductor component (usually a transistor) placed in parallel with
the load. Here it works
much like the overflow drain on a bathtub—if you fill the tub to the overflow drain, it
begins to siphon off water
to keep the tub from overflowing. With the shunt regulator, if the TEG is producing
power that would normally
take the load voltage beyond the desired operating point, the regulator will begin to
conduct. By drawing
additional current, it can drop more voltage across the internal resistance of the TEG
module.
Shunt
Load Regulator
INT
R
R
No-load
voltage
Figure 16
Thermoelectric Power Generation Technology FAQ 1 0
The shunt regulator can provide two types of regulation—line and load. In providing
line regulation, the shunt
circuit conducts current as necessary to hold the load voltage constant in the face of
changes in no-load voltage.
It can also provide load regulation—that is, altering its conductive properties to
compensate for changes in load
resistance. If the load requires more current, the shunt regulator conducts less; if the
load requires less current,
the shunt regulator conducts more.
There is one “fly in the ointment”, however. The shunt regulator can only provide
regulation as long as it is
drawing current. Once the circuit load resistance draws all of the available current at
the desired voltage, the
regulator can no longer respond to further reductions in either load resistance or no-
load voltage. If you design
your system to this worst case, however, you should not have any problem with
regulation.
It is important to note here, that the use of a regulator—either shunt or series—will
decrease the powerconversion
efficiency of the system (the power dissipated in the regulator is generated output
that is not
delivered to the load). This is simply part of the price of regulation, however.
29. Why not just control the heat source to get a regulated output?
This would be a terrific way to go if it was practical. With this approach, if a generator
was using excessive
energy, you could just cut down on the input power (e.g., fuel flow). The problem is
that this would typically
be done with an electrically-controlled valve. Unfortunately, the power demands of
these valves and the
support circuitry to control them, would consume most (and perhaps all) of the
generator’s potential output. An
electronic regulator is usually more cost-effective.
30. Is there a limit to how many devices I can use in a TEG system?
Theoretically, no, although at a certain point, it can get a little clumsy mechanically.
31. Are there any special considerations which apply to clamping a device?
The biggest challenge is finding an effective way to maintain fairly even compression
across all of the modules
employed. In doing this, you must provide enough compression screws (i.e., the
fasteners which draw the hot
and cold sides toward the module) to assure a good thermal interface, yet avoid any
distortion of the metal
stock (heat sinks, burner boxes, etc.). If the expanse is too great, the metal can
easily bow and compromise the
thermal interface (and mechanical integrity) in the process. Also, when assembling
the generator, it is important
to bring the compression up gradually on these screws so that modules are not
crushed from the collapsed
fulcrum effect {see “Mechanical Clamping Method”, in the Introduction to
Thermoelectrics}.
32. Do I need to be concerned with the rate of temperature change when
powering up my system?
You will certainly enhance the longevity of your system if you ramp it up (or down)
slowly. Very rapid changes
in temperature can cause damage to the module due to differential rates of thermal
expansion and contraction.
While some thermal variation is expected in a TEG system and the devices must
withstand this, it is still prudent
to minimize these dynamics whenever possible.
33. How can I determine what my heat sink needs are?
This is probably best done by using rule of thumb and inferences. Based on the
expected ⊗T, you can make
a guess on the system’s power conversion efficiency in the range of 3-5%. Then take
your worst case (i.e.,
greatest) electrical load in watts and divide it by the estimated efficiency. This will
approximate the number of
watts that must be dissipated in the sink. Next, determine a targeted temperature
difference between the hot side
and cold-side sink and divide it by the total number of watts that will be dissipated in
the sink. For example, if
we suspected an efficiency of 4% in an application delivering 5 W to a load, with a
temperature difference of
100° C between the heat sink and hot-side of the burner box, we could determine the
required heat sink/(fan)
resistance as follows:
Thermal Resistance ( )( ) ( ) ( ) CW
P
T Efficiency
Efficiency
P
T
LL
0.8 /
5
= ⊗ = 100 0.04 = °
 

 

= ⊗ 04
Thermoelectric Power Generation Technology FAQ 1
34. Would it be helpful to include a fan on the “cold side”?
With any heat sink, proper air movement will dramatically enhance the performance.
Thus, if you can provide
active ventilation to your cold-side sink, you should be able to improve your ⊗T. The
big question is whether
the improvement in performance will be more than enough to offset the power
demands of the fan. Sometimes
it will be feasible, sometimes not.
35. Do I have to insulate between the hot and cold sides of the system?
If you can find an insulating material which can withstand your highest temperatures,
it will improve the powerconversion
efficiency of the system.
36. How can I measure no-load voltage directly
when the system is connected to an electrical load?
Simply put, you can’t. The only way to measure the no-load voltage, is by
disconnecting the load and
measuring the voltage which results.
37. What happens if I design a system
which provides more electrical power than I need?
If the system always provides surplus power, try decreasing your ⊗T. This can be
done by either decreasing
input power (if possible) or by using a less efficient heat sink on the cold side. On the
other hand, if your power
surplus occurs on an intermittent basis and your load cannot withstand the
variability, you may need to manage
the excess with a shunt regulator.
38. Does Z-Max® offer any advantages over other thermoelectric power
generators?
Yes. With our unique, patented, hybrid metallurgy, Z-Max® devices yield power
outputs which are unsurpassed
in the industry.
39. Are these modules tested before shipment?
All Tellurex TEG’s are subjected to quality testing before shipment.
40. Can I check a TEG with an ohmmeter?
No. The vast majority of ohmmeters apply DC to the resistance being measured. This
will generate a Seebeck
voltage which makes the measurement inaccurate. To measure internal resistance
directly, you must use a meter
which applies a true AC waveform (i.e., one which varies in polarity every half-cycle).
With the continual
polarity reversals in this type of meter, little or no Seebeck voltage will be created
(although some self-heating
may result if you hold it under test for a prolonged period). To get this kind of
operation, you will probably
have to locate a high-quality LCR meter.
41. Can I use an ohmmeter which applies a pulsed DC?
No. This, too, will result in Seebeck voltages which make measurement inaccurate.
The meter must apply a
waveform which reverses polarity every half cycle.
42. How can I order?
Go to www.tellurex.com and click on the ordering page.
43. How can I get more info?
Call 231-947-0110.

1
Short title:- Thermoelectricity
Long Title:- Investigating the performance of a semiconductor, thermoelectric, heat
pump (TEC)
Weights:- 3 if all suggested experiment are done, including the modeling (step 3).
Scientific purpose:- From a series of experiments on a given thermoelectric cooling
(TEC)
device, determine values for the 3 device parameters in the usual physical model for
its steady
state performance. Use the model to predict optimum operating conditions for the
device and
estimate the greatest temperature drop it can provide (with zero heat input).
Heuristic purpose:- 1) To familiarize students with techniques for observing and
modelling the
performance of a thermodynamic system. 2) To observe heat production, heat flow,
and the
Seebeck and Peltier effects, and to understand their relationships with the first and
second laws
of thermodynamics.
_____________________________________________________________________________
Background:-
A TEC (Thermo-Electric Cooler) is an electrically driven, solid-state, heat pump in the
form of a
plate (Figure 1). In use, the input and output sides of the plate usually are thermally
connected to
different parts of the surrounding environment with thick pieces of aluminum metal.
Aluminum
conducts heat so readily that these so called "thermal reservoirs" will have an almost
uniform
temperature.
Figure 1. A semiconductor thermoelectric cooling module (TEC) (left). Internally, it consists of
alternating pairs of metal-semiconductor-metal thermocouple junctions (right). (From
www.peltierinfo.
com)
A TEC is an array of metal-to-semiconductor thermocouple junctions. If one side of
the TEC is at
a different temperature than the other, a potential difference will be observed
between its leads
(called the Seebeck voltage, Vs ). And if a suitable dc electrical current is passed
through the
TEC, the temperature of one side of the plate will drop with respect to the
temperature of the
other side. In steady operation, a TEC can suck heat from a cool environment on the
input side
and discharge heat (a greater amount of heat) to a warmer environment on the other
side.
Such a device is called a "heat pump" because it causes thermal energy to be moved
from a
cooler temperature to a warmer one, rather than the natural case of heat flowing
from higher to
lower temperature. There is, of course a penalty for frustrating nature's normal
habits, and this is
2
a necessary input of electrical power. Heat pumps are just the reverse of heat
engines. In a
classic heat engine, some amount of heat is input at high temperature, and a fraction
of that
amount is output at lower temperature, while the difference between the input
output heat
energies appears as mechanical (or some other kind of) work. In a heat pump or
refrigerator,
work energy is input to the device, which causes an amount of thermal energy to be
taken into
the device at some low temperature and a larger amount (the sum of the input heat
and electrical
energies) to be output at a higher temperature.
The “second law of thermodynamics” states that entropy (the heat energy
transported divided by
the absolute temperature at which it moves) must increase or (at best) remain
constant in the
process. The ideal case (no entropy change) is called a “reversible process”. Normal
heat flow
(heat energy moving from higher to lower temperature) always increases entropy.
Physical basis:-
The Seebeck and Peltier effects that a TEC exploits arise from the free (or partially
free) electrons
that allow charge to move through a conductive solid. The charge carriers also
transport heat
very effectively; this being in addition to whatever heat flows through the crystalline
frame of the
solid by lattice vibrations (called phonons). Especially in situations where different
conductive
media are in contact with one another, couplings arise between heat flow and current
flow, and
temperature difference and potential difference. A temperature difference can
generate a
potential difference (Seebeck effect), and a current flow can generate a heat flow
(Peltier effect),
and vice versa. The two effects are reciprocal; not independent. More information is
provided in a
separate document.
Apparatus:-
Figure 2 is a diagram of the refrigeration device (TEC) and its surroundings. The TEC
is
sandwiched between an aluminum block and an aluminum heat sink. These two
"thermal
reservoirs" (at temperatures Tin and Tout, units K) pass heat into and out of the TEC
through its
surfaces at rates Pin and Pout (positive flow is defined to be the direction from input
reservoir to
output, units J/s = W).
Figure 2. Schematic diagram of how a TEC can be used to cool an object. In our experiment,
heat is supplied electrically to the input reservoir by resistive dissipation and heat is removed
from
the output reservoir by a forced flow of air. (Illustration from www.peltier-info.com.)
The TEC consists of multiple pairs of metal-semiconductor-metal junctions in which
the
semiconductor type is alternated. The thermocouple pairs are connected electrically
in series
and thermally in parallel (Figures 1, 2 and 3).
3
Figure 3 (Right) TEC with one face removed:- bismuth telluride bricks, doped with impurities to
be p or n semiconductors. (Left) copper connection strips and top surface. (Illustration from
www.peltier-info.com).
Although the materials connecting the surfaces of the TEC between the junctions are
p and n
semiconductors, the device contains only metal to semiconductor junctions. Thus, a
TEC does
not exhibit the non-linear voltage versus current characteristic of a semiconductor
diode (which
requires a p-n junction). It acts like a battery (the thermocouple voltage) in series
with a small
ohmic resistance; and a potential difference of only a few volts applied across the
TEC can cause
an ampere level current to flow through it. The device itself is symmetrical; so the
direction of
current flow determines the direction of heat flow from one surface of the TEC to the
other.
Experimental set-up:-
In our experiment ( Figures 4, 5), heat is supplied at a controlled rate to the "input"
thermal
reservoir, by passing an ac electric current through a pair of resistors attached to it
(total
resistance Rh = 2.0 Ω; Pin = IhVh = Ih
2 Rh). The input reservoir is only poorly connected to the air
environment, so the heat flow into the TEC (Pin ) will be closely equal to the amount of
resistive
heating of the reservoir. (Of course, if the input reservoir is at a very different
temperature than
ambient air, it will receive or lose heat at a small rate, but this can be neglected, at
least initially).
The total flow of electrical power into the TEC is Pd = Vd Id where, respectively, Vd and
Id are the
total voltage across the device and the current flowing through it. This power is used
two ways,
as electrical work (Vs Id) done to overcome the TEC’s Seebeck (thermocouple) voltage
Vs, and as
ohmic power dissipation which heats the TEC (Id
2 Rd). The first part drives the Peltier heat pump;
the second part heats the TEC, and is an undesired side effect.
The output thermal reservoir is a fan-cooled, finned, aluminum heat sink which
dissipates Pout into
the air environment. To do this, the heat sink's temperature (Tout) must rise a few
degrees above
the ambient air temperature (T0).
The input reservoir gets (nearly) all its heat from the heating resistors (Rh). Thus Pin =
R h Ih
2 if the
temperatures of the TEC and its thermal reservoirs are not changing. Conservation of
energy
requires that Pout = Pin + Pd.
The temperature at the input and output of the TEC are monitored in the input and
output thermal
reservoirs. Locations of the electrical connection points and the temperature
monitoring points in
the apparatus are shown in Figure 5, along with all necessary circuitry.
4
Figure 5 Connection points on the experiment assembly.
Figure 6 Phenomenological model of TEC behaviour.
5
Phenomenological physical model of a TEC:-
The aim here is to use basic macroscopic physical principles to inter-relate different
aspects of
the TEC's performance. Included in the modelling equations are some unknown
parameters that
depend on the TEC's construction and its materials. If the values of these parameters
can be
estimated by a few experimental measurements, performance in other
circumstances can be
predicted. The standard phenomenological model of a TEC contains only three
elements:- 1) An
ideal Peltier (Carnot) reversible heat pump which is just an ideal thermocouple
working in
reverse, 2) Ohmic resistive heating (in addition to the Peltier effect) that occurs in the
TEC
semiconductor materials when current is passed through it. 3) Normal (non
electronic) heat flow
from the hot output port of the TEC to the cooler input port, such as occurs when the
TEC is
disconnected electrically. Figure 6 shows this schematically, along with the relevant
equations.
Remember that the model applies to steady state operation of the TEC. The TEC (and
especially
the thermal reservoirs attached to it) possess significant heat capacity. When input-
output
conditions are altered, it will take some minutes for the reservoirs and the TEC parts
to heat up or
cool down, and for a new steady state to be established.
Experimental procedures:-
Conceptually, the apparatus allows the TEC to be operated in three ways:-
1) with an approximately fixed current (Id ~ 5 A) supplied from the built in power
supply.
2) not operated, but with terminals open circuited (Id = 0).
3) not driven, but with the TEC terminals short circuited. However, this 3rd state is
hard to
achieve in practice, because a path needs to have very low (milliohm) resistance to
be effectively
a short circuit (zero resistance). The resistance of plug-in leads is often higher than
this. Also, it
may preclude insertion of an ammeter, because the meter's resistance may be too
large. If you
wish to do such an experiment, connections must be made with thick wire firmly
clamped into the
screw terminals of the apparatus, and current measurement must be obtained by
reading (with an
electronic micro-voltmeter) the very low voltage produced across a very low value
resistor often
called a "current shunt" or "current sense resistor" (10 mΩ or 1 mΩ).
Heat can be supplied to the input thermal reservoir at any desired rate by adjusting
the 60 Hz
voltage output from a "Variac" variable transformer which is applied (via a separate
step-down
transformer) to the pair of series connected 1.0 ohm resistors. Power inputs in the
range of about
0 - 20 W are feasible.
When the TEC and its thermal reservoirs are in steady operation, the thermal inertia
of the
system allows turning off the TEC current Id for a few seconds without much
disruption of its
temperature state. Thus, one can record the total Seebeck thermocouple voltage
being generated
in the TEC simply by breaking the current supply circuit for a few seconds and
measuring the
voltage across the TEC carefully.
Suggested experiments:-
1) Passive heat flow experiment (TEC open circuited, heat sink fan on).
For several values of Pin from 0 to about 5 W, measure the steady state
temperatures of the
input and output reservoirs, and the ambient air,. Graph the temperature differences
Tin –Tout and
Tout – T0 as a function of Pin ( = Ph), and try to estimate the thermal conductance Kd of
the TEC
(between the two reservoirs) and Khs between the output heat sink and the ambient
air.
6
2) Cooling experiments, (Active TEC)
2a ) After letting the system cool off after experiment 1, operate the TEC with no
heat input to the
input reservoir for about 15 minutes, while monitoring the voltage across the TEC
and the current
supplied to it, as well as the input and output temperatures. Draw a graph of the
temperatures
and the electrical power input Pd versus time. It will be hard to establish a completely
steady
state, but after a few minutes, changes in state should be slow enough that the
temperatures and
power input Pd can effectively be measured nearly simultaneously.
2b) After about 15 minutes of cooling (in 2a), or once the temperature jump across
the TEC has
reached a few tens of degrees, make a series of observations where a flow of heat is
supplied to
the TEC’s input reservoir. Also, as well as monitoring Pd (Vd with Id on), make
observations of the
TEC’s Seebeck voltage (Vd with Id off) for a few seconds every minute or so.
In steps lasting a few minutes each (so a steady state is approximated) apply
increasing amounts
of power to the resistors on the input reservoir (0 to ~10W). Monitor the reservoir
temperatures,
Pin and Pd (with Id on) and also Vd (with Id = 0), and plot graphs of the reservoir
temperature
differences, Pd and Vd versus Pin.
3 Modelling (use of MatLab or Mathematica is recommended.)
3a) After examining the modelling equations given in Figure 6, select results from 1
and 2b to
estimate values for the model parameters of the apparatus and the TEC.
3b) Use the TEC model to predict the results of 2a (maximum cooling with no heat
input).
3c) Decide what would be the optimum drive current for the device if it is to be
operated (say)
with an output temperature of 35C and an input temperature not exceeding 5 C.
When operated
at this current, what is the maximum rate at which heat can be supplied the input
reservoir.
(without it warming above 5 C)?
3e) What is the lowest input temperature that can achieved with this device (using
the optimum
input current), if absolutely no heat is admitted to the input reservoir?
Possible Additional experiments:-
The model described above does not take account of thermal conductances /
resistances
between the heat sink and the environment, or the input reservoir and the
environment, or in the
TEC between the semiconductor elements and the surface of the TEC. Can you
estimate these
from observational data you can obtain or by looking up propertied of materials in
handbooks and
create a more complete model of the TEC. How would you make a dynamic model
(one that
would predict warm-up /cool-down rates).
References:-
http://en.wikipedia.org/wiki/Peltier_cooler
good place to start
http://www.ferrotec.com/usa/thermoelectric/ref/3ref2.htm
http://www.peltier-info.com/info.html
and links at bottom of Introduction page
http://www.tf.uni-kiel.de/matwis/amat/elmat_en/kap_2/backbone/r2_3_3.html
(Theory of Seebeck and Peltier effects)

Thermoelectrics

Thermoelectrics and how it works

The Thermoelectric Peltier effect is the most direct way to


utilize electricity to pump heat. Electric current forces the
matter to approach a higher energy state and the heat is
absorbed (cooling).
Simply by reversing the direction of the current a heating
effect is created.

Power Generator Seebeck (1822)

Some of the heat input is converted to electric current


(work), as the higher energy matter (black dots) releases
energy and cools to a lower energy state (white dots).
The net work is proportional to the temperature difference
and Seebeck coefficient.
Heat Pump (Refrigerator) Peltier (1834)

Electrical current (work input) forces the matter to


approach a higher energy state (black dots) and heat is
absorbed (cooling).

The energy is released (heating) as the matter


approaches a lower energy state (white dots). The net
cooling or heating effect is pro-portional to the electric
current and Peltier coefficient.

In simple terms, the thermoelectric Peltier effect is the


most direct way to utilise electricity to pump heat.

Thermoelectric modules

The material used at working temperatures up to 150¡C is


normally bismuth telluride, doped to obtain p (positive)
and n (negative) semi-conducting properties.

A number of pn-couples, thermally parallel and


electrically in series, are sandwiched between ceramic
plates.

The maximum temperature differential (Delta Tmax)


between the cold and the warm side of a Supercool
single stage module is up to 74¡C at warm side
temperatures of 25¡C or Delta T = 84¡C at a warm side
temperature of 50¡C. By increasing the number of stages
in a multistage arrangement, you also increase maximum
Delta T.
Thermo Life™
Thermo Life is a thermoelectric source generating power in concert with a
thermoelectric principle that converts heat flow directly into electrical energy.

Wherever a temperature difference can be used, it


is a self-generating energy source with an output
of approximately 10 µW at voltages in the volt
range.

Based on a temperature differential gradient, heat


always seeks a cold surface. The Thermo Life
generator when attached to a cold surface collects
heat until it warms up to the same temperature as
the heat moving to it.

Thermoelectrics is the science and technology associated with thermoelectric


converters, that is, the generation of electrical power by the Seebeck effect and
refrigeration by the Peltier effect. Thermoelectric generators are being used in
increasing numbers to provide electrical power in medical, military, and deep
space applications where combinations of their desirable properties outweigh
their relatively high cost and low generating efficiency. In recent years there also
has been an increase in the requirement for thermoelectric coolers (Peltier
devices) for use in infrared detectors and in optical communications.
Thermoelectrics
Rama Venkatasubramanian Ph.D., Senior Program Director, rama@rti.org

Thermoelectric materials are of interest for applications as heat pumps and power
generators. The performance of thermoelectric devices is quantified by a figure of merit,
ZT, where Z is a measure of a material's thermoelectric properties and T is the absolute
temperature. A material with a figure of merit of around unity was first reported over four
decades ago, but since then — despite investigation of various approaches — there has
been only modest progress in finding materials with enhanced ZT values at room
temperature. We have observed a ZT ~2.4 in our Bi2Te3/Sb2Te3 superlattice devices.The
enhancement is achieved by controlling the transport of phonons and electrons in the
superlattices — more specifically, phonon-blocking, electron-transmitting superlattices.

New thermoelectric materials could make a life-or-death


difference for the next generation of microprocessors,
enable fiber optic switches 100 times faster than current
technology, and even turn an automobile's waste heat
into electricity to run the air conditioner.

University of Birmingham

Physics and Space Research Degree

Course 3 Laboratory

Experiment 1: Cooled Detectors in Spacecraft

ABSTRACT
Spacecraft in Earth orbit are in a hostile thermal environment,
alternately baking in the Sunlit half of the orbit and then
freezing in the Earth’s shadow. Many detectors such as CCD’s,
need to be cooled to low temperatures, to reduce dark counts.
This experiment investigates the performance of Peltier heat
pumps, and methods of heat dissipation into space by cold
fingers and passive thermal radiators.

Theory

Transfer of Heat by Conduction

The thermal conductivity, k is defined as the rate at which heat is conducted through a
material and is expressed in units of W m-1 K-1

The quantity of heat Q flowing in Watts is given by

Q = k A (T1 –T2) / d

Examples of the thermal conductivity of various materials are given below.


Material k W m-1 K-1

Copper 390
Aluminium 237
Stainless Steel 16
Glass 0.93
Water 0.6
Nylon 0.25
Expanded polystyrene 0.03
Air 0.025

These values are a function of temperature!

The specific heat of a material is the amount of heat required to raise the temperature of
1 kg of the material by 1 Kelvin and has units J kg-1 K-1

Examples of specific heats of various materials are given below.

Material Specific Heat J kg-1 K-1

Water 4180
Aluminium 889.7
Iron 449

Specific heats are also a function of temperature!

Transfer of Heat by Radiation

The radiation emitted by a black body is given by the expression

Power (in watts) P =Aσ T4

Where A is the surface area of the black body, σ is the Stefan’s constant and T is the
absolute temperature.
For a body which is not black then we have to take into account the emissivity ε of the
surface, so we obtain:

P =A ε σ T4

In the space environment, these are the only methods of heat transfer that we need to
consider, as there can be no convection in the vacuum of space. However in the
laboratory we may also need to consider convection.
Transfer of Heat by Convection

The transfer of heat by convection obeys Newton’s Law of Cooling, which states that the
rate of transfer of heat is proportional to the excess temperature ΔT. The amount of heat
transferred Q is given by

Q=h A ΔT

Where, A is the surface area, ΔT is the temperature difference, and h is a convection


coefficient, which depends on many factors including whether the convection is natural
or forced, (with a fan) and also whether the flow is streamlined or turbulent.

The Peltier Device

The Peltier device is a series of P-N junctions through which and electric current is
passed. One side of the junction becomes hot and the other cold. The first part of this
laboratory experiment is for you to understand the physics of a Peltier device and to write
a paragraph in your laboratory book, explaining the detailed physics of how such a device
operates. This information can be obtained from a textbook or from the WWW, but must
be written in your own words.

The diagram below shows a schematic heat engine, which can represent the Peltier
device.

H Heat is pumped from the cold reservoir to the hot


reservoir by means of the work done by the electrical
power W. Clearly Q1 = W + Q2

Ex

Experimental Equipment

A Peltier device, RS Components type 618-730 is attached by thermal glue to one end of
a cold finger. At the other end of the cold finger is a copper heat sink, cooled by water
(RS type169-4404). The cold finger is made of aluminium alloy and is 40.0mm in
diameter. A series of ‘J-type’ thermocouples is inserted along the cold finger and on the
hot and cold sides of the Peltier device, as well as in the water inlet and exit tubes. The
thermocouples are at 30.0mm spacing along the cold finger.

Cold Finger

Water in
T1 T2 T3 T4 T5
Water out

Peltier Heat Pump

Water Heat Exchanger

You are provided with a DC Power supply to operate the Peltier Heat Pump. The
maximum temperature of the hot junction must not exceed 70C, nor must the DC current
exceed a maximum of 8.0A. Operation outside these values will destroy the device.
Whenever the Peltier device is powered, the cold finger must be cooled by a flow of
water from a constant head apparatus. This device will produce a constant flow of water
through the heat sink. The actual flow can be adjusted by changing the heights of the inlet
and exit pipes. A flow of between 50ml and 150ml per minute should be sufficient.
A Pico Technology 8 channel thermocouple data logger connected to a PC is used to
measure and record temperatures in the experiment. Manuals for the data logger and
specification sheets for the Peltier device can be found in Appendix 1 of this experiment
manual.

Experiment 1: Characteristics of a Peltier Heat Pump

Determine the optimum performance of the Peltier heat pump, by measuring the
temperatures of the hot and cold junctions as a function of the input electrical power.
In any thermal experiment it is essential that the whole system be in thermal equilibrium,
otherwise erroneous results will be obtained. This will depend on the thermal time
constant of the system. It would be useful, therefore to estimate the thermal time
constant, before making any measurements. How did your estimate of the thermal time
constant compare with the time it actually took for thermal equilibrium to be established?
That is when there is no significant change in any of the thermocouple temperatures.

Plot a graph of cold junction temperature against the voltage and determine the optimum
voltage. Record the water flow and the temperature difference in the water.

The efficiency of the heat pump is the amount of heat, QC, pumped from the cold to the
hot junction by the electrical power (I*V)
This is often called the coefficient of performance, COP = QC / (I*V). The following
graph shows how COP is a function of I / IMAX and ΔT /Δ T MAX

I / I MAX

Graph of coefficient of performance (COP) as a function of I / I MAX


And ΔT / ΔT MAX

You have the values of I MAX and Δ T MAX from the data sheets, so you should be able to
determine the efficiency (COP) of the device, and knowing the electrical power be able to
estimate the heat pumped QC . This will then allow you to calculate the total heat pumped
into the cold finger, which will be useful in the next part of the experiment.

Experiment 2: Thermal Conductivity of the Cold Finger

The heat flow Q down the cold finger is given by the expression

Q = k A (T1 –T2) / d

Where k is the thermal conductivity, A is the area of cross section of the cold finger and
(T1 –T2) / d is the temperature gradient along the cold finger.

Set up the experiment so that the Peltier hot junction is at about 50C and the rate of water
cooling is about 100ml per minute. Allow the experiment to come into thermal
equilibrium and then measure the temperature along the cold finger with the
thermocouples T1 to T5 to determine the temperature gradient. Plot a graph of the
temperature along the cold finger using a spreadsheet.
How linear is the gradient? What can this tell you ?

You can then determine the amount of heat, Q, flowing down the cold finger by the
temperature increase of the water. Alternatively you can use a measurement of the
electrical power and the COP to determine the heat flow down the cold finger.

Estimate the thermal conductivity k of the cold finger and consider the errors in the value
you obtain. The alloy of the cold finger is aluminium type BS6082. Compare the value
you have obtained for k with published values.

Project Section

You are provided with a second cold finger made of the same material and with an
identical Peltier heat pump in the same insulated container. Instead of having a water-
cooled heat sink, the cold finger has a mounting flange to attach passive radiators.
Engineering drawings of the cold fingers and radiators can be found in Appendix II.

In space experiments, the heat produced by a Peltier heat pump, used to cool a detector, is
radiated into deep space by a passive radiator. This is often coated with a material of high
emissivity ε and is shielded from the Sun.

Investigate the performance of the two passive radiators, which have been provided.
You are provided with a supply of thermocouples and data logger.

What is the lowest detector temperature possible with each of the radiators?

How does convection affect the performance of the radiators? Consider repeating your
experiments in a vacuum chamber if this is available.

Investigate how the surface finish and coating affects the emissivity of the passive
radiators and therefore their performance.

Investigate how the shape of passive radiators affects their performance.

Look at the actual designs of passive radiators in spacecraft and consider making and
testing a new radiator on the cold finger.

How might both of the cold fingers be improved to improve their performance in a
laboratory environment?

A CCD detector and cold finger are available, which you can interface to the Peltier
device with the passive radiator. Investigate the performance of the CCD, when mounted
on the passive radiator.
Appendices

Appendix I RS Data Sheet – Peltier effect heat pumps

Appendix II Melcor SealTEC Sealed Thermoelectric Coolers data sheet

Appendix III PICO TECH Thermocouple Manual

You might also like