ArticlePDF Available

Abstract and Figures

More than 50% of the energy released through combustion in the internal combustion engine (ICE) is rejected to the environment. Recovering only a part of this energy can significantly improve the overall use of resources and the economic efficiency of road transport. One solution to recoup a part of this otherwise wasted thermal energy is to use thermoelectric generator (TEG) modules for the conversion of heat directly into electricity. To aid in development of this technology, this effort covers the derivation of a respectively simple steady‐state Simulink model that can be utilized to estimate and optimize TEG system performance for ICEs. The model was validated against experimental data found in literature utilizing water cooling for the cold side. Overall, relatively good agreement was found with the maximum error in generated power around 10%. Following, it was investigated whether air can be used as a cooling medium. It was established that, at the same temperature as the water (18.4 °C), a flow velocity of 13.1 m/s (or 47.2 km/h) is required to achieve a similar cold junction temperature and power output. Subsequently using the model with air cooling, the performance of a TEG installed on a heavy‐duty vehicle traveling at 50, 80, 90, and 120 km/h under different ambient temperatures was analyzed. It was determined that both a lower temperature and a higher flow velocity can improve power output. A further increase of the power output requires a larger temperature gradient across the module, which can be achieved by a higher heat input on the hot side.
Content may be subject to copyright.
applied
sciences
Article
Simulink Model of a Thermoelectric Generator for Vehicle
Waste Heat Recovery
Nicolae Vlad Burnete 1, Florin Mariasiu 1, * , Dan Moldovanu 1and Christopher Depcik 2


Citation: Burnete, N.V.; Mariasiu, F.;
Moldovanu, D.; Depcik, C. Simulink
Model of a Thermoelectric Generator
for Vehicle Waste Heat Recovery.
Appl. Sci. 2021,11, 1340. https://
doi.org/10.3390/app11031340
Academic Editor: Alberto Benato
Received: 18 January 2021
Accepted: 29 January 2021
Published: 2 February 2021
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2021 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
1Automotive Engineering and Transports Department, Technical University of Cluj-Napoca,
400641 Cluj-Napoca, Romania; nicolae.vlad.burnete@auto.utcluj.ro (N.V.B.);
dan.moldovanu@auto.utcluj.ro (D.M.)
2Department of Mechanical Engineering, University of Kansas, Lawrence, KS 66045, USA; depcik@ku.edu
*Correspondence: florin.mariasiu@auto.utcluj.ro
Featured Application: The work presented in this study can be used in the design and develop-
ment process of thermoelectric generators for vehicle waste heat recovery.
Abstract:
More than 50% of the energy released through combustion in the internal combustion
engine (ICE) is rejected to the environment. Recovering only a part of this energy can significantly
improve the overall use of resources and the economic efficiency of road transport. One solution
to recoup a part of this otherwise wasted thermal energy is to use thermoelectric generator (TEG)
modules for the conversion of heat directly into electricity. To aid in development of this technology,
this effort covers the derivation of a respectively simple steady-state Simulink model that can be
utilized to estimate and optimize TEG system performance for ICEs. The model was validated against
experimental data found in literature utilizing water cooling for the cold side. Overall, relatively
good agreement was found with the maximum error in generated power around 10%. Following,
it was investigated whether air can be used as a cooling medium. It was established that, at the
same temperature as the water (18.4
C), a flow velocity of 13.1 m/s (or 47.2 km/h) is required to
achieve a similar cold junction temperature and power output. Subsequently using the model with
air cooling, the performance of a TEG installed on a heavy-duty vehicle traveling at 50, 80, 90, and
120 km/h under different ambient temperatures was analyzed. It was determined that both a lower
temperature and a higher flow velocity can improve power output. A further increase of the power
output requires a larger temperature gradient across the module, which can be achieved by a higher
heat input on the hot side.
Keywords:
thermoelectric generator; thermoelectric module; waste heat recovery; internal combustion
engine; steady-state model
1. Introduction
With the advent of the internal combustion engine (ICE) in the late 19th century and
its continued development during the 20th century, humankind has benefited greatly
from its use for transportation. In fact, the 20th century can be defined in one word:
“Mobility”; local, regional, national, and international mobility of goods, people, and
ideas. However, owing to its construction and general operation, the ICE manages to only
partially transform fuel chemical energy into mechanical work. This is illustrated by the
thermal efficiency of modern ICEs that generally only reaches 40–45%. This respectively
low value of energy conversion directly influences the economic efficiency of the road
transport process. For example, a 40 t freight vehicle with a payload of 25 t operating with
an ICE has an overall transport process efficiency equal to 17–18% [
1
]. Furthermore, since
transportation has a significant influence in both national and global economies, increasing
transport process efficiency can immediately reduce operating costs while additionally
helping to decrease greenhouse gas emissions.
Appl. Sci. 2021,11, 1340. https://doi.org/10.3390/app11031340 https://www.mdpi.com/journal/applsci
Appl. Sci. 2021,11, 1340 2 of 33
Most of the thermal energy of the ICE (approx. 30%) is lost through the high tempera-
ture combustion exhaust gases that are discharged directly into the atmosphere. Therefore,
one solution to increase overall efficiency is to recover, at least a part of this significant heat
loss to the environment. For this purpose, research has targeted the design, development,
and implementation of thermoelectric generator modules for directly converting heat into
electricity, which can then be used in hybrid power systems to increase vehicle efficiency.
Thermoelectric devices are environmentally friendly and have numerous other advantages
including the lack of moving parts, the conversion of thermal energy directly into electrical
energy, high reliability, as well as low maintenance requirements. The main disadvantage
of such devices is their low efficiency owing to the small dimensionless figure of merit (
ZT
)
values currently found in commercial devices. However, there are promising solutions for
achieving higher ZT values even at low to mid temperatures (<150 C) [2].
The research efforts focused on identifying the potential for converting thermal energy
into electricity, through thermoelectric generator (TEG) systems of different configurations
and sizes, installed on the engine’s exhaust path show that it is possible to obtain an
electrical power output of approximately 200–1000 W, depending on vehicle type and
operating conditions [
3
6
]. However, research has also shown that the vehicular use of
TEG systems depends on driving conditions with negative effects found when conditions
have a high changing frequency [7].
The development of numerical models investigating the efficient use of TEG systems
centers mainly on: Operating temperatures, optimizing geometric dimensions [
8
], different
materials in TEG construction [
9
], and design optimization of hot and cold side heat
exchangers [
10
]. Furthermore, TEG system design focuses on the energetic efficiency
of conductive and convective system types [
11
]. In this area, Fan et al. [
12
] proposed a
numerical model to determine the optimum values for the thermocouple leg lengths and
cross-sectional areas. Their analysis showed that, for given thermal boundary conditions,
the maximum output power can be improved by optimizing the leg length or cross-sectional
area. Other efforts by Yu et al. [
13
] developed a mathematical model to investigate the
effects of different vehicle engine operating regimes (e.g., cold-start and different loads)
on TEG performance. The major conclusion of their research was that TEG performance
is influenced more by vehicle speed than by ambient temperature. Wang et al. [
14
] used
a three-dimensional computational fluid dynamics simulation to analyze power output,
temperature distribution, and pressure drop of a TEG system with different internal
structures of the heat exchanger. Their model considered the coverage and placement of
the thermoelectric modules on the heat exchanger surface. The primary outcome was that
increasing the number of modules may eventually saturate the total power output. Du
et al. [
15
] investigated the design of a cooling channel for a thermoelectric module using a
CFD software program (Fluent). Their results showed that in general, the output power is
higher when using liquid as cooling fluid compared to air.
The analysis of this literature shows that there is a wide variety of numerical models
for TEG simulation with different types of sophistication. However, most models are
usually only briefly presented, with little detailing of the underlying equations and of the
implementation. Therefore, the aim of this study is to develop and validate a respectively
simple steady-state Simulink model that can be utilized to estimate and optimize TEG sys-
tem performance. The model is intended to aid in the design and development of thermal
energy recovery systems for ICEs, but it can be applied in other areas as well. Employing
the described model early in the design process can facilitate whether thermal energy
recovery is feasible for the TEG system under specific operating conditions. Furthermore,
the proposed Simulink model can be integrated into existing software solutions for the
analysis of different energy sources and powertrains.
2. Thermoelectric Model Equations
A thermoelectric couple (TC) consists of p-type and n-type semiconductor legs that are
connected thermally in parallel and electrically in series (Figure 1). When heat is applied
Appl. Sci. 2021,11, 1340 3 of 33
to produce electric current, the TC operates as a generator. If, however, current is applied
at the terminals, then the TC operates as a heat pump. For this effort, the focus is on
thermoelectric generation and the use of the model as a heat pump is left to future work.
Figure 1. Thermoelectric couple operating as a generator.
If the two junctions of the TC are subjected to different temperatures (
T = T
h
T
c
[K], where T
h
[K]—hot side temperature, T
c
—cold side temperature), a potential difference,
proportional to the temperature variation, will appear (V
OC,TC
[V]—open circuit voltage,
VLoad,TC [V]—potential difference across a load connected to the TC):
VOC,TC =αTC ·(ThTc)(1)
where αTC [V/K] is the relative Seebeck coefficient:
αTC =αpαn. (2)
If the TC is connected to an external load with a resistance R
Load,TC
[
], a current flow
will appear (I [A]). Consequently, the potential difference VLoad,TC can be expressed as:
VLoad,TC =I·RLoad,TC. (3)
In addition, there is an internal electrical resistance of the TC, R
TC
[
], which is defined
based on the characteristics of the two legs: Electrical resistivity (
ρp
[
m]—electrical
resistivity of the p-type leg,
ρn
[
m]—electrical resistivity of the n-type leg), length (L
p
[m]—length of the p-type leg, L
n
[m]—length of the n-type leg), and cross-sectional area (A
p
[m2]—cross-section area of the p-type leg, An[m2]—cross-section area of the n-type leg).
RTC =ρp
Lp
Ap
+ρn
Ln
An(4)
Knowing that the relation between VLoad,TC and VOC,TC is
VLoad,TC =VOC,TC IRTC (5)
which results in a derivation of the current:
Appl. Sci. 2021,11, 1340 4 of 33
I=αTC(ThTc)
RTC +RLoad,TC
. (6)
This flow of current leads to three additional effects: Peltier heating or cooling, Joule
heating, and the Thomson effect. The latter is relatively small and is usually neglected to
simplify the analysis, as is the case in this study.
If current flows through the junction, heat must be continuously added or rejected to
maintain a constant temperature. Furthermore, this heat flow is proportional to the current
flow and changes direction if the current is reversed. This is called the Peltier effect and is
defined by the Peltier coefficient πTC,junction [J/A]:
πTC,junction =
.
Qjunction
I(7)
where
.
Qjunction
[W] is the rate of heat absorption/rejection at the hot/cold junction. Us-
ing Kelvin relationships, the Peltier coefficient can be expressed based on the Seebeck
coefficient, which is respectively easier to measure:
πTC,junction =αTCTjunction (8)
where T
junction
[K] is the temperature of the hot side and cold side junction. Thus, the rate
of heat absorption/rejection at this junction can be expressed as:
.
Qjunction =αTCITjunction . (9)
In addition to Joule heating, there is another irreversible phenomenon that accom-
panies the reversible thermoelectric effect, namely the conduction of heat. In a TC, heat
conduction causes a transfer of heat from the hot side to the cold side as a means of
achieving thermal equilibrium. Considering the thermal conductivities of the two legs (k
p
[W/mK]—thermal conductivity of the p-type leg, k
n
[W/mK]—thermal conductivity of
the n-type leg), one can define the thermal conductance (KTC [W/K]) of the TC:
KTC =kp
Ap
Lp
+knAn
Ln. (10)
Joule heating,
.
QJoule,h/c
[W], is the process where electrical energy is converted into
thermal energy as current flows through a resistance. For a TC, Joule heating decreases the
rate of heat absorption at the hot junction and increases the rate of heat transfer to the cold
junction, in equal shares (i.e., half of the heat generated by the Joule effect passes to the hot
side and the other half to the cold side) [16]:
.
QJoule,h/c =1
2I2RTC. (11)
Neglecting the Thomson effect, by applying the first law of thermodynamics at the
hot junction of a TC, the rate of heat absorption, .
Qh,TC [W], is:
.
Qh,TC =αTCITh1
2I2RTC +KTC(ThTc)(12)
whereas the rate of heat rejection, .
Qc,TC [W], at the cold side is:
.
Qc,TC =αTCITc+1
2I2RTC +KTC(ThTc). (13)
Overall, the power delivered to the load, Pel,TC [W], can be calculated in two ways:
Pel,TC =.
Qh,TC
.
Qc,TC (14)
or as:
Appl. Sci. 2021,11, 1340 5 of 33
Pel,TC =I2RLoad,TC. (15)
The thermoelectric performance of the material is assessed using the figure of merit [1/K]:
Z=α2
ρk(16)
which, for a TC becomes
ZTC =α2
TC
hqρpkp+pρnkni2. (17)
This figure of merit, like the thermoelectric parameters, is temperature dependent. How-
ever, often, average values are assumed with errors found to be below 10% of the true
value [17].
The conversion (or thermal) efficiency of the TC is the ratio of the power output to
the heat input at the hot side. To simplify the final expression, the following notations are
employed for the load resistance ratio (m [-]) and Carnot efficiency (ηC[-]), respectively,
m=RLoad,TC
RTC
(18)
ηC=1Tc
Th
. (19)
Thus, the conversion efficiency is:
ηth =ηCm
(1+m)1
2ηC+1
2ZT (1+m)21+Tc
Th. (20)
Expressions of maximum parameters (current, voltage, power, and efficiency) are also
introduced here due to their importance in determining TEG parameters, which will be
shown later. The maximum voltage occurs at open circuit when I = 0:
VOC,TC,max =αTC(ThTc). (21)
The maximum current, however, occurs at short circuit, where R
Load,TC
= 0; thus,
leading to the following expression:
Imax =αTC(ThTc)
RTC
. (22)
To obtain the maximum power, it can be shown that the load resistance must equal
the internal resistance (thus: m = 1), which leads to the following expression:
Pel,max =α2
TC(ThTc)2
4RTC
(23)
For the maximum efficiency, the conversion efficiency is differentiated with respect to m
and set to 0. This leads to the following expression for the load resistance ratio:
m=p1+ZT (24)
where
ZT
[-] is the dimensionless figure of merit (
T=
1
/
2
·(Th+Tc)
[K] is the average
temperature across the TC). Consequently, the maximum conversion efficiency becomes:
ηth,max =1Tc
Th·p1+ZT 1
p1+ZT +Tc
Th
. (25)
Of additional high importance is the efficiency at maximum power (
ηmp
), or, in other
words, at matching load resistance (m = 1). Based on Equation (20), ηmp is:
Appl. Sci. 2021,11, 1340 6 of 33
ηmp =ηC
21
2ηC+2
ZT 1+Tc
Th. (26)
The output of a single TC is generally small; e.g., well below 1 [W]. Therefore, many
TCs are connected to form a thermoelectric module (TEM) (Figure 2); thus, increasing
the power output. At this point it is worth recalling that a TEM can operate both as
a generator (often designated with TEG in literature) and/or as a heat pump (usually
TEC—thermoelectric cooler). In the present study, the TEM operates as a generator, but the
notation of TEM is used to avoid confusion with the acronym TEG that involves the overall
system including heat exchangers (shown later). The output of a TEM can be expressed
based on the number of thermocouples (n
TC
[-]) and the parameters of the TC, except for
electrical current that is independent of the number of thermocouples. Thus, the total rate
of heat absorption/rejection at the hot and cold sides are, respectively:
.
Qh,TEM =nTCαTC ITh1
2I2RTC +KTC(ThTc)(27)
.
Qc,TEM =nTCαTC ITc+1
2I2RTC +KTC(ThTc). (28)
Appl. Sci. 2021, 11, 1340 7 of 36
Figure 2. Thermoelectric module operating as a generator (based on [18]).
3. TEM Properties Identification Model
Often TEM material properties are not available from the manufacturer and must be
determined, which requires specialized equipment [19]. Furthermore, it is not always
necessary to have high model accuracy throughout the entire operating range. For
example, when estimating the performance of a TEM system in the preliminary design
phase, it is helpful to have a respectively simple and physically based model that can be
used to run parametric studies. To solve this issue, Lee et al. [20] and Zhang [21] proposed
methods that use maximum parameter values, usually provided by manufacturers, to
estimate the electrical resistivity, Seebeck coefficient, thermal conductivity, and the figure
of merit. The authors call these effective parameters since, as compared to intrinsic
material properties, they are determined at the system (TEM) level; thus, they account for
the various losses caused by the manufacturing process (i.e., thermal and electrical contact
resistances), the temperature dependency (due to the Thomson effect), as well as imperfect
insulation (heat losses through radiation and convection). The basic idea is that, ideally,
the maximum current, voltage, power, and efficiency are functions of the material
properties, the thermocouple geometry (area, length), and the temperatures of the two
junctions (T
h
, T
c
).
3.1. Equations and Simulink Model
For the current study, the maximum parameters used to determine the outcomes are
current, efficiency, and power (only three of the four maximum parameters presented in
Equations (21)–(23) and (25) are required). By knowing the leg area (A = A
p
= A
n
) and
length (L = L
p
= L
n
), the effective Seebeck coefficient and the effective electrical resistivity
of the TC can be determined from Equations (22) and (23):



ρ=
el,m ax
TC 2
TC max
A
4P
L
nI (30)
()
α=
el, max
TC
TC max h c
4P
nI T T . (31)
To determine the effective thermal conductivity, an additional parameter is required,
namely the effective figure of merit which can be obtained from Equation (25):
Figure 2. Thermoelectric module operating as a generator (based on [18]).
Using Equations (14), (27) and (28), the total power output of the module can
be calculated
:
Pel,TEM =nTChαTC I(ThTc)I2RTCi. (29)
3. TEM Properties Identification Model
Often TEM material properties are not available from the manufacturer and must
be determined, which requires specialized equipment [
19
]. Furthermore, it is not always
necessary to have high model accuracy throughout the entire operating range. For example,
when estimating the performance of a TEM system in the preliminary design phase, it
is helpful to have a respectively simple and physically based model that can be used
to run parametric studies. To solve this issue, Lee et al. [
20
] and Zhang [
21
] proposed
methods that use maximum parameter values, usually provided by manufacturers, to
estimate the electrical resistivity, Seebeck coefficient, thermal conductivity, and the figure of
merit. The authors call these effective parameters since, as compared to intrinsic material
properties, they are determined at the system (TEM) level; thus, they account for the
various losses caused by the manufacturing process (i.e., thermal and electrical contact
resistances), the temperature dependency (due to the Thomson effect), as well as imperfect
insulation (heat losses through radiation and convection). The basic idea is that, ideally, the
Appl. Sci. 2021,11, 1340 7 of 33
maximum current, voltage, power, and efficiency are functions of the material properties,
the thermocouple geometry (area, length), and the temperatures of the two junctions
(Th, Tc).
3.1. Equations and Simulink Model
For the current study, the maximum parameters used to determine the outcomes are
current, efficiency, and power (only three of the four maximum parameters presented in
Equations (21)–(23) and (25) are required). By knowing the leg area (A = A
p
= A
n
) and
length (L = L
p
= L
n
), the effective Seebeck coefficient and the effective electrical resistivity
of the TC can be determined from Equations (22) and (23):
ρ
TC =4A
LPel,max
nTCI2
max
(30)
α
TC =4Pel,max
nTCImax (ThTc). (31)
To determine the effective thermal conductivity, an additional parameter is required,
namely the effective figure of merit which can be obtained from Equation (25):
Z
TC =1
T
1+ηmax
ηC
Tc
Th
1ηmax
ηC!2
1
. (32)
Based on Equation (16), the effective thermal conductivity can be expressed as:
k
TC =α
TC2
Z
TCρ
TC
. (33)
3.2. Model Validation
The model has been validated for two commercially available thermoelectric modules
which are presented in Table 1. This data has been used in the Simulink model presented
in Figure 3to determine the effective parameters.
1
3
Figure 3. Simulink model for determining effective properties (values for TGM-127-1.4-2.5).
Appl. Sci. 2021,11, 1340 8 of 33
Table 1. Thermoelectric parameters of commercially available thermoelectric modules.
TEM nTC [-] Pel,max [W] Imax [A] ηmax [%] Ap/n [mm2]Lp/n [mm] Th [C] Tc [C]
TGM-199-1.4-0.8 [22] 199 11.40 5.10 4.3 1.96 0.8 200 30
TGM-127-1.4-2.5 [23] 127 4.46 2.37 5.5 1.96 2.5 200 30
The effective parameters obtained with the model presented in Figure 3were used to
calculate the voltage, current, and power at matched load resistance conditions, as well as
the power, current, voltage, and efficiency at the maximum temperature gradient for various
values of the load resistance. The results show that there is a relatively good agreement be-
tween the measured and calculated values (Figures 4and 5). The major differences appearing
here can be attributed to the temperature dependency of the electrical resistivity, Seebeck
coefficient, and thermal conductivity, which is not captured in the proposed model. Despite
these differences, the predicted power output of the module is in relatively good agreement
with the measured values. A possible reason for this result is that the maximum power was
chosen as an input parameter in deference to the maximum voltage.
Figure 4.
Manufacturer data compared to calculated data at matched load resistance conditions for TGM-199-1.4-0.8:
(
a
) Module power; (
b
) module current; (
c
) module voltage; (
d
) module power and voltage with respect to current; and
(e) module efficiency.
Appl. Sci. 2021,11, 1340 9 of 33
Figure 5.
Manufacturer data compared to calculated data at matched load resistance conditions for TGM-127-1.4-2.5:
(
a
) Module power; (
b
) module current; (
c
) module voltage; (
d
) module power and voltage with respect to current; and
(e) module efficiency.
4. TEG Model for Exhaust Gas Waste Heat Recovery
For the current study, the system is analyzed under steady-state conditions. The do-
main is discretized along the flow direction in control volumes (CVs) having the same
length as a TEM (Figure 6). Initially, only one control volume is considered for the calcula-
tion; however, a greater domain can be built by adding more control volumes with each
taking input data from the previous CV. The changes in fluid properties with temperature
are considered separately for each control volume and are defined at the mean CV temper-
ature (average of inlet and outlet temperatures). For simplicity, thermoelectric parameters
are considered constant, but the model can be readily updated with equations for these
parameters, either from the literature or from the manufacturer.
Thermoelectric modules are often combined with heat exchangers (HX) to form ther-
moelectric generators (TEG), which are used to recover otherwise wasted thermal energy
from the exhaust gases of ICEs. The fins of the HX allow for an increased heat transfer
surface area, subsequently enhancing the amount of heat extracted through convection
from the hot gases. This heat is then transferred through conduction to the TEMs. An
Appl. Sci. 2021,11, 1340 10 of 33
example of such a TEG with all relevant dimensions of the finned heat exchanger (L
HX
[m]—heat exchanger length, W
HX
[m]—heat exchanger width, H
HX
[m]—heat exchanger
height, z
HX,fin-spacing
[m]—heat exchanger fin spacing, t
HX,fin
[m]—heat exchanger fin thick-
ness), as well as the positioning of the TEMs is presented in Figure 6. If the heat exchanger
is covered with TEMs on both sides, the CV must be modified accordingly (Figure 7).
Figure 6. Hot side heat exchanger.
Figure 7. Side view of module section.
4.1. Equations and Simulink Model
The layering of the TEG, which will be used for calculating the amount of heat trans-
ferred to the hot side and from the cold side, is presented in Figure 8. The TEG is composed
of two heat exchangers (hot and cold side), a thermal grease used to eliminate surface
Appl. Sci. 2021,11, 1340 11 of 33
imperfections in the interface area, and two ceramic plates acting as electric insulators. The
electrical conductor height is relatively small and, since it has a respectively high thermal
conductivity and low electrical resistivity, it was neglected in the present model.
Figure 8. Integration of thermoelectric module in a thermoelectric generator.
As a first step, all relevant areas for the heat transfer are calculated starting from the
area available for heat transfer in one TC (Aht,TC [m2]):
Aht,TC =Ap+An. (34)
For a TEM with n
TC
thermocouples, the area available for heat transfer (A
ht,TEM
[m
2
])
will be:
Aht,TEM =nTCAp+An. (35)
Thus, for a TEG with n
TEM
[-] TEMs, the total area of the HX base in a control volume
(Aht,CV,base-TEM [m2]) available for heat transfer will be:
Aht,CV,baseTEM =nTEMnTC Ap+An. (36)
The HX base area for one control volume (Aht,CV,base [m2]) is defined as:
Aht,CV,base =WHx ·LCV (37)
where L
CV
is the length of the CV. The part of the HX base area that is not covered by TEMs,
is considered to be insulated.
To calculate the amount of heat extracted from the hot gases, the total area (A
ht,CV,total
)
available for heat exchange with the hot gases in a control volume must be determined:
Aht,CV,total =nfinh2(tHx,fin +LCV )HHx +zHx,finspacingLCVi. (38)
Note that the equations presented are for a HX installed on the hot side, but the
equations can also be used for an HX on the cold side.
To show the fraction of the HX base surface that is used for heat transfer to the TCs, a
HX surface usage coefficient is proposed:
fHx,Area =Aht,CV,baseTEM
Aht,CV,base
. (39)
When only the flow velocity (V
gas
[m/s]) is available, the crossflow area of the HX is
required to compute the mass flow:
AHx,cross =nfinzHx,finspacing HHx (40)
consequently, the gas mass flow rate will be
Appl. Sci. 2021,11, 1340 12 of 33
.
mgas =ρgasVgas AHx,cross. (41)
If, however, the mass flow is available, then Equation (41) can be used to compute the
gas flow velocity.
The next step is to determine the forced convection heat transfer coefficient between
the gases and the HX surface. This is respectively difficult since the flow ranges from
developing to fully developed. On this topic, Teertstra et al. [
24
] proposed and validated a
model that uses a composite solution based on the limiting cases of flow between isothermal
parallel plates. In the transitional flow region, an average Nusselt number is calculated as
a function of the heat exchanger geometry and fluid flow velocity. To use this model, two
conditions must be met: Air must be contained within the channels (this is achieved since
a closed HX is considered) and the fin spacing must be considerably smaller than the fin
height (high aspect ratio). The reduced Reynolds number has an experimental range of
0.2 Re200 [25] and is calculated as:
Re
Hx,channel =VgaszHx,finspacing
µgas
zHx,finspacing
LHx
. (42)
The proposed equation for the Nusselt number is:
NuHx,channel =
Re
Hx,channelPrgas
2!3
+
0.664qRe
Hx,channelPr 1
3v
u
u
t1+3.65
qRe
Hx,channel
3
1
3
. (43)
If the flow is fully turbulent and Re * exceeds its range limits, then the Nusselt number
can be calculated using the Gnielinski relation [26]:
NuHx,channel =
fG
8(ReHx,channel 1000)Prgas
1+12.7qfG
83
qPr2
gas 1(44)
where Re
HX,channel
is the channel Reynolds number and f
G
is the friction factor in turbulent
flow in smooth tubes determined from the first Petukhov equation [27]:
ReHx,channel =VgaszHx,finspacing
µgas
(45)
fG=1
(0.79 ln ReHx,channel 1.64)2. (46)
Knowing the Nusselt number, the gas-fin heat transfer coefficient can then be calcu-
lated with the following expression:
hgasfin =kgas
zHx,finspacing
NuHx,channel. (47)
To calculate the efficiency of the heat exchanger, as well as the thermal resistance, the
efficiency of a single fin must be determined [
28
], along with the thermal resistances of all
other intermediate layers. A single fin efficiency is calculated based on the m
fin
parameter
and the fin height, which, in this case is the same as the heat exchanger height:
mfin =shgasfin ·2(LHx +tHx,fin)
kHxLHx tHx,fin
(48)
ηfin =tanh(mfinHHx )
mfinHHx
. (49)
The next step is to calculate the overall effectiveness of the fin (
ηo,fin
[-]), for which the
area of a single fin (Afin [m2]) and the number of fins (nfin [-]) must be known:
Appl. Sci. 2021,11, 1340 13 of 33
Afin =2HHx(LHx +tHx,fin )(50)
ηo,fin =1nfinAfin
Aht,Hx,base
(1ηfin). (51)
Figure 9b shows the thermal resistance network used to calculate the hot (T
h
[K]) and
cold side (Tc[K]) junction temperatures along with the heat exchanger efficiency. To keep
the model simple, and since it represents under 2% of the total heat transferred from the
hot side [
29
], radiation losses are neglected. The resistances in Figure 9are defined for
all layers presented in Figure 8based on their thickness (t
layer-name
), thermal conductivity
(k
layer-name
), and surface area (A
layer-name
). The equivalent gas-fin heat transfer resistance,
and the thermal resistances of the base, grease, and ceramic are calculated according to
Equations (52)–(55):
Rth,fin,eq =1
ηo,finhgasfin Aht,Hx,total
(52)
Rth,Hx,base =tHx,base
kHx,baseAht,Hx,base
(53)
Rth,grease =tgrease
kgreaseAgrease (54)
Rth,ceramic =tceramic
kceramicAceramic
(55)
Figure 9.
(
a
) Thermoelectric generator (TEG) temperatures of interest; (
b
) thermal resistance network for a CV; (
c
) simplified
thermal resistance network for calculating Tc; and (d) simplified thermal resistance network for calculating Th.
The equivalent thermal resistance of the TEM is calculated based on the hot and cold
junction temperatures and the heat flow to the cold side:
Rth,TEM =ThTc
.
Qc
. (56)
Appl. Sci. 2021,11, 1340 14 of 33
Similarly, an equivalent load resistance is defined based on the hot junction temper-
ature and the temperature on the cold side (T
,c
= 1/2
(T
,c,in
+ T
,c,out
) [K], where
T
,c,in
and T
,c,out
are the temperature of the gas at the cold side heat exchanger inlet and
outlet respectively):
Rth,Load,eq =ThT,c
Pel,TEM
. (57)
Based on Equations (52)–(55), an equivalent thermal resistance of the hot side can
be determined:
Rth,h,eq =Rth,fin,eq +Rth,Hx,base +Rth,grease +Rth,ceramic. (58)
Similarly, the equivalent thermal resistance of the cold side is:
Rth,c,eq =Rth,ceramic +Rth,grease +Rth,Hx,base +Rth,fin,eq. (59)
Often, an aluminum block is used to act as a thermal spreader that is placed between
the HX and the TEM as in Figure 10. In this case, the thermal resistance of the base is
replaced by the thermal resistance of the aluminum block:
Rth,Al,block =tAl,block
kAl,blockAAl,block
. (60)
Figure 10. Integration of thermoelectric module in a thermoelectric generator.
It is important to note that, in such cases, the HX is usually also made of aluminum
which allows for a simplification. Namely, the addition of the HX base thickness to the
aluminum block thickness. They can, however, be introduced as separate blocks. In the
present study, the former solution has been adopted.
For the case presented in Figure 10, the heat exchanger efficiency is defined as:
ηHx =1
ηo,fin
+2hgasfinAht,Hx,total tgrease
kgreaseAgrease
+hgasfinAht,Hx,total tAl
kAlAAl . (61)
To compute the junction temperatures, one more thermal resistance is needed, namely
the equivalent thermal resistance of the out (output and heat rejection) side:
Appl. Sci. 2021,11, 1340 15 of 33
Rth,out,eq = 1
Rth,Load,eq
+1
Rth,TC +Rth,c,eq !1
. (62)
Based on these resistances, the hot and cold side junction temperatures can be ex-
pressed with respect to T
,h
(T
,h
= 1/2
(T
,h,in
+ T
,h,out
) [K], where T
,h,in
and T
,h,out
are the temperature of the gas at the hot side heat exchanger inlet and outlet respectively)
and T,c:
Th=Rth,out,eqT,h +Rth,h,eq T,c
Rth,out,eq +Rth,h,eq
(63)
Tc=Rth,c,eqTh+Rth,TC T,c
Rth,c,eq +Rth,TC
. (64)
The amount of heat extracted from the hot gases can be determined from the CV inlet
and outlet enthalpy represented by the heat capacity times the difference in temperature:
Qgas,Hx =.
mgascp,gas (T,h,in T,h,out). (65)
However, in this effort, the outlet temperature is an output value. Neglecting the heat
losses through radiation Qgas,HX = Qh, therefore, the outlet temperature is:
T,h,out =T,h,in Qh
.
mgascp,gas
. (66)
When installing a heat exchanger on the exhaust path of an ICE, there is also the issue
of the increased backpressure that negatively influences engine performance and emissions.
Consequently, the pressure drop must be estimated. Here, there are three effects that need
to be considered: Viscous drag effects, flow expansion, and flow contraction. For a CV, the
pressure loss due to this viscous drag effect is:
Ploss,Hx =
CV
fCVlength
Dhydr,channel
ρgas
V2
channel
2(67)
where f is the friction factor defined using the Churchill equation [30]:
f=8"8
Rechannel 12
+1
(A+B)1,5 #1
12
. (68)
The Churchill equation was chosen because it covers both developing and fully
developed internal flows. The Reynolds number is calculated using Equation (45), while A
and B are:
A=(2.457 ln"7
Rechannel 0.9
+0.27 εsurface
Dhydr,channel #)16
(69)
B=37350
Rechannel 16
. (70)
To calculate A, two additional parameters are required, namely the surface rough-
ness (
εsurface
), which can be taken from literature and the channel hydraulic diameter
(Dhydr,channel), which is calculated from the channel cross-sectional area and perimeter:
Dhydr,channel =2zHx,finspacingHHx
zHx,finspacing +HHx
. (71)
For small gas density changes along the length of the heat exchanger, in case of n
CV
control volumes (CVlength = LHX/nCV), Equation (67) can be simplified to:
Appl. Sci. 2021,11, 1340 16 of 33
Ploss,Hx =fLHx
Dhydr,channel
ρgas
V2
channel
2. (72)
Changes in the pipe cross-section area along the flow direction also lead to pressure
losses by causing the flow to either expand or contract. To estimate these losses, the
Borda-Carnot correlations are used [
31
]. These correlations require knowledge about cross-
sectional areas, as well as the gas density and velocity. For a TEG installed on the exhaust
pipe of an ICE (Figure 11), these cross-sectional areas are: Exhaust pipe upstream of the
HX (A
inlet-pipe,cross
[m
2
]), HX inlet (A
HX,inlet,cross
[m
2
]) and outlet (A
HX,outlet,cross
[m
2
]), as
well as the exhaust pipe downstream of the HX (A
outlet-pipe,cross
[m
2
]). Usually, TEG heat
exchangers have a constant cross-section; thus, A
HX,inlet,cross
= A
HX,outlet,cross
= A
HX,cross
.
The inlet and outlet pipe cross-sectional areas are:
Ainletpipe,cross =π
d2
inletpipe
4(73)
Aoutletpipe,cross =π
d2
outletpipe
4. (74)
Figure 11. Areas used for calculating the pressure loss due to flow contraction and expansion.
The pressure loss due to flow expansion occurs before the HX, thus leading to
the expression:
Ploss,exp =ρgas
Ainletpipe,cross
AHx,cross 1Ainletpipe,cross
AHx,cross V2
gas (75)
whereas the pressure loss due to flow contraction occurs after exiting the HX:
Appl. Sci. 2021,11, 1340 17 of 33
Ploss,contr =1
2ρgas1
µcontraction
12Ainletpipe,cross
AHx,cross 2
V2
gas (76)
with µcontraction as the contraction coefficient:
µcontraction =0.63 +0.37 AHx,cross
Ainletpipe,cross !3
. (77)
It must be noted that the pressure losses due to flow expansion and contraction are
calculated only for the HX inlet and outlet CVs, respectively.
4.2. Solution Method
The thermal circuit temperatures must be solved in an iterative manner due to the
dependence of the thermal resistances on the terminal temperatures and thermoelectric
properties. Therefore, before beginning the iterative calculation process, several temper-
atures must be estimated (i.e., T
h
, T
c
, T
,h,out
, and T
,c,out
). Using input data and these
estimated temperatures, the thermal resistances together with the other parameters are eval-
uated. Then, the values of T
h
, T
c
, T
,h,out
, and T
,c,out
are reevaluated and the convergence
is checked based on the relative error value:
Errorrel =
TiTi+1
Ti
(78)
If Error
rel
is below a set value (10
6
for this study), the calculation stops. If this is not
the case, the initially guessed values are updated with the current results and the process
starts again. The solution methodology flowchart for each CV is presented in Figure 12.
Figure 12. Solution methodology flow chart.
5. Results and Discussions
The model (Figure 13) was validated based on tests performed by Fagehi et al. [
32
]. In
their work, Fagehi et al. [
32
] used a liquid (water) heat exchanger on the cold side which
demanded a series of changes to the cold side heat exchanger model (not included in this
study, but available upon request).
The comparison between the simulated and measured values is presented in
Figure 14
.
Fagehi et al. [
32
] measured the junction temperatures, current, and voltage directly, while
they calculated the power using two methods. The power designated with P
el_Temp_I
is
calculated using the junction temperatures, electric current, and the ideal thermoelectric
equations; whereas, P
el_V_I
was determined basedon the measured voltage and current. Sim-
ilar to the results of Fagehi et al. [
32
] the simulation results were closer to the power values
Appl. Sci. 2021,11, 1340 18 of 33
determined using the junction temperatures, electric current, and the ideal thermoelectric
equations. The average and matched load resistance errors for the junction temperatures,
power output, current, voltage, and efficiency are listed in Table 2. The possible causes of
differences are: Measurement accuracy, the use of ideal equations, effective material prop-
erties, as well as constant values for both the thermoelectric properties and gas properties.
In addition, this effort neglects the Thomson coefficient, radiative heat transfer, and the
thermal spread resistance. Another possible source of error is the extraction of data from
the graphs presented in literature by using a digital tool. Despite all the uncertainties, there
is a reasonable agreement between the calculated and the measured values.
2
13
Figure 13.
Simulink model with two heat exchangers (red—hot side and blue—cold side), aluminum
blocks for thermal spread (light grey), the TEM between, and the thermal resistance network on
right side.
Table 2. Comparison of calculated and measured values at peak measured power.
Parameter Th[C] Tc[C] Pel [W] I [A] V [V] ηth [-]
Average error [%] 0.15 0.33 12.73/1.77 6.82 0.52 0.93
Error at matched load resistance [%] 0.11 0.45 10.46/3.50 2.99 0.38 1.8
As previously mentioned, Fagehi et al. [
32
] used a liquid (water) heat exchanger on
the cold side. The same cold side temperature can be achieved by using air as the cooling
fluid, but this poses an issue due to the very low thermal conductivity of air. This can be
resolved by increasing the flow velocity and, to test this theory, the water cold side heat
exchanger model (used only for the validation) was replaced with an air heat exchanger
(similar to the one used on the hot side). Then, the air flow velocity on the cold side was
adjusted to achieve a T
c
value as close as possible to that obtained with the validated model
(Figure 14). The required flow velocity was determined to be 13.1 m/s (or 47.2 km/h)
and appears feasible for highway trucks and buses running at a constant speed. Here, the
main advantages of utilizing air would be the lower weight and space requirements of the
system, as well as a reduced complexity. Furthermore, a proper design of the flow passage
to the cold side can allow for high air velocities; thus, leading to even greater temperature
gradients than in the case of the liquid coolant solution. The main disadvantages of this
solution are that the temperature gradient and, consequently, the power output depends
Appl. Sci. 2021,11, 1340 19 of 33
on vehicle speed and outside temperature (still less than the coolant temperature). Also,
this solution is not feasible for stationary engines.
Figure 14.
TEG model validation highlighting the difference between the simulated and measured values of (
a
) the electrical
power output, (
b
) hot and cold junction temperatures, (
c
) current, and (
d
) voltage for two different types of cold side heat
exchangers: Water (W_sim) and air (air_sim).
To test the possible output of such a TEG, 16 test cases (Table 3) were evaluated
with varying air temperatures (T
air
) and air flow velocities (V
air
). For the hot side, the
temperature of the exhaust gases was set at 250
C with a flow velocity of 19.1 m/s (
12 L
engine running at 1500 min
1
, with an exhaust pipe diameter of 100 mm). For brevity,
only the results for 50 and 90 km/h are presented here with the others are available upon
request. Regarding case notation, “V” stands for velocity and “T” for temperature. The
values following “V” and “T” are the velocity (in km/h) and temperature (in
C) values for
that case (e.g., V50T-5 = velocity of 50 km/h and temperature of
5
C). For this study it
was assumed that the hot and cold side fluid temperatures and flow velocities are uniform
and constant, but also that there is a uniform temperature distribution on the TEG and
TEM surface. Furthermore, the simulation was performed for only one TEM. Consequently,
when considering the desired number of TEMs this will result in an overprediction of the
power output.
Appl. Sci. 2021,11, 1340 20 of 33
Table 3. Test cases for cold side heat exchanger using air as cooling fluid.
Case Tair,-5 [C] Tair,10 [C] Tair,20 [C] Tair,30 [C]
Vair,50 = 13.9 [m/s]
510 20 30
Vair,80 = 22.2 [m/s]
Vair,90 = 25.0 [m/s]
Vair,120 = 33.3 [m/s]
Table 4and Figures 15 and 16 show the results for the two different velocities 50 and
90 km/h. Analysis of the results finds that both the higher air flow velocity on the cold
side and a lower ambient temperature increase the output of the TEG. Decreasing the air
temperature raises the temperature gradient and the power output, but also reduces the
temperature of the hot side junction to that below the maximum allowable value of the
TEM. This is because the hot side heat flow rate is constant. To further augment the output
of the TEG, it would be necessary to increase the heat flow rate on the hot side (either by
increasing the exhaust temperature or the flow velocity). Consequently, a high TEG output
requires not only a good HX, but also a suitable heat extraction strategy.
Table 4. Results for cold side heat exchanger using air as cooling fluid.
Parameters
Case
V50T-5 V50T10 V50T20 V50T30 V90T-5 V90T10 V90T20 V90T30
Pmax [W] 4.023 3.562 3.260 2.972 4.568 4.043 3.702 3.376
Th[K] 449.2 453.5 456.4 459.3 444.6 449.0 452.1 455.2
Tc[K] 347.6 357.6 364.7 371.7 336.3 347.1 354.6 362.0
T [K] 101.7 95.9 91.8 87.6 108.3 101.9 97.5 93.1
Figure 15.
(
a
) Module power output, (
b
) junction temperature, (
c
) module current, and (
d
) module voltage with respect to
the load resistance ratio for a velocity of 50 km/h at various temperatures (5, 10, 20, and 30 C).
Appl. Sci. 2021,11, 1340 21 of 33
Figure 16.
(
a
) Module power output, (
b
) junction temperature, (
c
) module current, and (
d
) module voltage with respect to
the load resistance ratio for a velocity of 90 km/h at various temperatures (5, 10, 20, and 30 C).
6. Conclusions
Considering the vast amounts of wasted thermal energy from internal combustion
engines, the study of thermoelectric power generation is necessary as a possibility to
increase overall energy efficiency. Thermoelectric devices are environmentally friendly
and have numerous other advantages including the lack of moving parts, high reliability,
the direct conversion of thermal energy into electrical energy, as well as low maintenance
requirements. However, the main disadvantage of such devices is their low efficiency
owing to the small ZT values currently found in commercial devices.
In the present study, two Simulink models are presented (see Appendix Afor detailed
Simulink blocks). The first model was created based on the work of Lee et al. [
20
] to estimate
thermoelectric parameters of a thermoelectric module. The validation showed that, even
though the temperature dependency of the thermoelectric parameters was neglected, there
is a relatively good agreement with experimental data.
Furthermore, a TEG model was proposed and validated for power output estimations
in the case of exhaust gas thermoelectric generators. Model validation was accomplished
using the results of Fagehi et al. [
32
]. For the validation, the cold side heat exchanger model
was adapted to match the cooling method of the authors with reasonable agreement. The
next step was to study the possibility of using air as the cooling fluid and to achieve the
same cold side temperature, the flow velocity of the cold air was adjusted (an air velocity
of 13.1 m/s showed a relatively good agreement).
After determining that similar results can be obtained with air as the cooling fluid, a
parametric study was performed to show the potential of such a system for ICE exhaust
coupling. The test conditions (used as input data for the cold side heat exchanger) consid-
Appl. Sci. 2021,11, 1340 22 of 33
ered a vehicle (with a respectively large engine capacity) traveling at different speeds (50,
80, 90, and 120 km/h) under dissimilar ambient air temperatures (
5, 10, 20, and 30
C).
The results showed that either a high air velocity or a low outside temperature leads to
greater TEG power output. For the analyzed cases, the peak power output was 4.568 W
for a temperature gradient of 108.3 K. It was also found that increasing the heat rejection
on the cold side requires a simultaneous growth in heat input at the hot side to facilitate
a favorable temperature gradient and, therefore, the power output of the TEG. Overall,
this demonstrates that it is feasible to install an air-cooled TEG on a heavy-duty vehicle
running at constant speed.
Author Contributions:
Conceptualization, N.V.B.; methodology, N.V.B.; software, N.V.B. and D.M.;
validation, F.M., C.D., and N.V.B.; formal analysis, N.V.B. and F.M.; investigation, N.V.B.; resources,
N.V.B.; data curation, N.V.B. and D.M.; writing—original draft preparation, N.V.B.; writing—review
and editing, N.V.B., F.M., and C.D.; visualization, N.V.B.; supervision, F.M.; project administration,
N.V.B. All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not Applicable.
Informed Consent Statement: Not Applicable.
Data Availability Statement: Not Applicable.
Acknowledgments:
This paper was supported by the Project POCU/380/6/13/123927 “Entrepreneurial
competences and excellence research in doctoral and postdoctoral programs-ANTREDOC”, project
co-funded by the European Social Fund.
Conflicts of Interest: The authors declare no conflict of interest.
Appendix A
Detailed Simulink blocks (Figures A1A14). Please note that the values that can be
seen in the figures were not necessarily used to validate the model. Also, the iterative
temperature calculation loop is not included here.
Heat exchanger–hot side (similar for cold side)
3
A1
Figure A1. Hot side heat exchanger block.
Appl. Sci. 2021,11, 1340 23 of 33
Figure A2. Hot side heat exchanger block—inputs.
Appl. Sci. 2021,11, 1340 24 of 33
Figure A3. Hot side heat exchanger block—HX heat transfer equations and outputs.
Appl. Sci. 2021,11, 1340 25 of 33
Figure A4. Hot side heat exchanger block—HX pressure drop equations and outputs.
Aluminum block—hot side (similar for cold side)
4
A5
Figure A5. Aluminum block.
Figure A6. Aluminum block—equations.
Appl. Sci. 2021,11, 1340 26 of 33
Thermoelectric module
5
A7
Figure A7. Thermoelectric module block.
Figure A8. Thermoelectric module block—inputs and outputs.
Appl. Sci. 2021,11, 1340 27 of 33
Figure A9. Thermoelectric module block—equations—part 1.
Appl. Sci. 2021,11, 1340 28 of 33
Figure A10. Thermoelectric module block—equations—part 2.
Appl. Sci. 2021,11, 1340 29 of 33
Thermal resistance network
Figure A11. Thermal resistance network block.
Appl. Sci. 2021,11, 1340 30 of 33
Figure A12. Thermal resistance network block—equations.
Appl. Sci. 2021,11, 1340 31 of 33
Figure A13. Effective properties model blocks (Figure 3)—equations.
Figure A14. Effective properties model blocks (Figure 3)—Detail of the effective material properties block (Figure A13).
Appl. Sci. 2021,11, 1340 32 of 33
References
1.
Hariram, A.; Koch, T.; Mårdberg, B.; Kyncl, J. A Study in Options to Improve Aerodynamic Profile of Heavy-Duty Vehicles in
Europe. Sustainability 2019,11, 5519. [CrossRef]
2.
Márquez-García, L.; Beltrán-Pitarch, B.; Powell, D.; Min, G.; García-Cañadas, J. Large Power Factor Improvement in a Novel
Solid–Liquid Thermoelectric Hybrid Device. ACS Appl. Energy Mater. 2018,1, 254–259. [CrossRef]
3.
Thacher, E.F.; Helenbrook, B.T.; Karri, M.A.; Richter, C.J. Testing of an automobile exhaust thermoelectric generator in a light
truck. Proc. Inst. Mech. Eng. Part D J. Automob. Eng. 2007,221, 95–107. [CrossRef]
4.
Liu, X.; Deng, Y.D.; Li, Z.; Su, C.Q. Performance analysis of a waste heat recovery thermoelectric generation system for automotive
application. Energy Convers. Manag. 2015,90, 121–127. [CrossRef]
5.
Yu, C.; Chau, K.T. Thermoelectric automotive waste heat energy recovery using maximum power point tracking.
Energy Convers. Manag.
2009,50, 1506–1512. [CrossRef]
6.
Karri, M.A.; Thacher, E.F.; Helenbrook, B.T. Exhaust energy conversion by thermoelectric generator: Two case studies.
Energy Convers. Manag. 2011,52, 1596–1611. [CrossRef]
7.
Yu, S.; Du, Q.; Diao, H.; Shu, G.; Jiao, K. Effect of vehicle driving conditions on the performance of thermoelectric generator.
Energy Convers. Manag. 2015,96, 363–376. [CrossRef]
8.
Ma, X.; Shu, G.; Tian, H.; Yang, H.; Chen, T. Optimization of length ratio in segmented thermoelectric generators for engine’s
waste heat recovery. Energy Procedia 2019,158, 583–588. [CrossRef]
9.
Deng, Y.D.; Liu, X.; Chen, S.; Tong, N.Q. Thermal Optimization of the Heat Exchanger in an Automotive Exhaust-Based
Thermoelectric Generator. J. Electron. Mater. 2013,42, 1634–1640. [CrossRef]
10.
Marvão, A.; Coelho, P.J.; Rodrigues, H.C. Optimization of a thermoelectric generator for heavy-duty vehicles.
Energy Convers. Manag.
2019,179, 178–191. [CrossRef]
11.
He, W.; Wang, S.; Zhao, Y.; Li, Y. Effects of heat transfer characteristics between fluid channels and thermoelectric modules on
optimal thermoelectric performance. Energy Convers. Manag. 2016,113, 201–208. [CrossRef]
12.
Fan, L.; Zhang, G.; Wang, R.; Jiao, K. A comprehensive and time-efficient model for determination of thermoelectric generator
length and cross-section area. Energy Convers. Manag. 2016,122, 85–94. [CrossRef]
13.
Yu, S.; Du, Q.; Diao, H.; Shu, G.; Jiao, K. Start-up modes of thermoelectric generator based on vehicle exhaust waste heat recovery.
Appl. Energy 2015,138, 276–290. [CrossRef]
14.
Wang, Y.; Li, S.; Zhang, Y.; Yang, X.; Deng, Y.; Su, C. The influence of inner topology of exhaust heat exchanger and thermoelectric
module distribution on the performance of automotive thermoelectric generator. Energy Convers. Manag.
2016
,126, 266–277. [CrossRef]
15.
Du, Q.; Diao, H.; Niu, Z.; Zhang, G.; Shu, G.; Jiao, K. Effect of cooling design on the characteristics and performance of
thermoelectric generator used for internal combustion engine. Energy Convers. Manag. 2015,101, 9–18. [CrossRef]
16. Goldsmid, H.J. Introduction to Thermoelectricity, 2nd ed.; Springer: Berlin/Heidelberg, Germany, 2016; ISBN 978-3-662-49255-0.
17. Rowe, D.M. Thermoelectrics Handbook—Macro to Nano; CRC/Taylor & Francis: Boca Raton, FL, USA, 2006.
18.
Kumar, P.M.; Jagadeesh-Babu, V.; Subramanian, A.; Bandla, A.; Thakor, N.; Ramakrishna, S.; Wei, H. The Design of a Thermoelec-
tric Generator and Its Medical Applications. Designs 2019,3, 22. [CrossRef]
19.
Wang, H.; McCarty, R.; Salvador, J.R.; Yamamoto, A.; König, J. Determination of Thermoelectric Module Efficiency: A Survey.
J. Electron. Mater. 2014,43, 2274–2286. [CrossRef]
20.
Lee, H.; Attar, A.M.; Weera, S.L. Performance Prediction of Commercial Thermoelectric Cooler Modules using the Effective
Material Properties. J. Electron. Mater. 2015,44, 2157–2165. [CrossRef]
21.
Zhang, H.Y. A general approach in evaluating and optimizing thermoelectric coolers. Int. J. Refrig.
2010
,33, 1187–1196. [CrossRef]
22.
KRYOTHERM. TGM-199-1.4-0.8. Available online: https://kryothermtec.com/assets/dir2attz/ru/TGM-199-1.4-0.8.pdf (ac-
cessed on 4 January 2021).
23.
KRYOTHERM. TGM-127-1.4-2.5. Available online: http://kryothermtec.com/assets/dir2attz/ru/TGM-127-1.4-2.5.pdf (accessed
on 4 January 2021).
24.
Teertstra, P.; Yovanovich, M.M.; Culham, J.R.; Lemczyk, T. Analytical forced convection modeling of plate fin heat sinks. In Pro-
ceedings of the Fifteenth Annual IEEE Semiconductor Thermal Measurement and Management Symposium (Cat. No.99CH36306),
San Diego, CA, USA, 9–11 March 1999; pp. 34–41.
25. Lee, H. Thermoelectrics: Design and Materials; John Wiley & Sons: Hoboken, NJ, USA, 2016.
26. Gnielinski, V. New equations for heat and mass transfer in turbulent pipe and channel flow. Int. Chem. Eng. 1976,16, 359–368.
27.
Petukhov, B.S. Heat Transfer and Friction in Turbulent Pipe Flow with Variable Physical Properties. In Advances in Heat Transfer; Hartnett, J.P.,
Irvine, T.F., Greene, G.A., Cho, Y.I., Eds.; Elsevier: Amsterdam, The Netherlands, 1970; Volume 6, pp. 503–564. ISBN 0065-2717.
28.
Cengel, Y.; Ghajar, A. Heat and Mass Transfer: Fundamentals and Applications, 6th ed.; McGraw-Hill Education: New York, NY, USA,
2020; ISBN 978-981-315-896-2.
29.
Kumar, S.; Heister, S.D.; Xu, X.; Salvador, J.R.; Meisner, G.P. Thermoelectric Generators for Automotive Waste Heat Recovery
Systems Part II: Parametric Evaluation and Topological Studies. J. Electron. Mater. 2013,42, 944–955. [CrossRef]
30.
Cengel, Y.; Cimbala, J. Fluid Mechanics Fundamentals and Applications, 4th ed.; McGraw-Hill Education: New York, NY, USA, 2017;
ISBN 978-1259696534.
31.
Bergman, T.; Adrienne, L.; Frank, I.; David, D. Fundamentals of Heat and Mass Transfer; John Wiley & Sons, Inc.:
New York, NY, USA
,
2011; ISBN 978-0470-50197-9.
32.
Fagehi, H.; Attar, A.; Lee, H. Optimal Design of an Automotive Exhaust Thermoelectric Generator. J. Electron. Mater.
2018
,47,
3983–3995. [CrossRef]
... Thermoelectric p-n couple unit resistance, r T C ( ), is expressed as [18,19]: ...
... Contribution to the Ohm law of the TEG output voltage U T E G (V) will offer here the value of the output current I T E G (A), through relation [18]: ...
... The relation between the TEG output voltage (in the presence of a load) and open circuit voltage U OC (R L = 0) is defined as follows [18]: ...
Article
Full-text available
Nowadays, high-performance electronic devices release large amounts of heat throughout their operation. Thermoelectric passive cooling systems can offer a cost-efficient and straightforward solution for the thermal management of such devices based on a passive heatsink and a thermoelectric generator (TEG). TEG can transform a part of the heat generated by the electronic chip into electricity for the power supply of microelectronic systems. The thermoelectric system investigated here contains a 12 W Positive Temperature Coefficient (PTC) heating element placed on a 4 × 4 cm 2 aluminium plate, a commercial SP1848-27145 TEG device and a finned plate heatsink recycled from a desktop PC. Two types of tests were performed here. An energy efficiency test was initiated at different load resistance values R L between 1 and 40 by directly connecting TEG output to a direct current (DC) electronic load. The second test was developed for electrical power generation at 3.3 V-5 V by integrating a DC-DC boost converter between the TEG output and the electronic load set at R L of 600-900. It was registered a maximum energy conversion efficiency of 1.17 % for R L = 2 , close to the internal resistance value of the TEG module. The second test revealed the generation of output electrical powers with minimum values of 15-17.5 mW at a temperature difference between TEG plates of about 20 K.
... The above equation makes it abundantly evident that a higher ZT is equivalent to enhanced power production performance; hence, this ZT will dictate the efficiency with which the TEG conversion occurs (Burnete et al. 2021). In addition to this ZT, the temperature differential between the two sides of the TEG is also a significant factor in determining the TEG's overall net effectiveness, as the following equation demonstrates: ...
Article
Full-text available
Green hydrogen is a buzzword in today's renewable energy applications research domain. This paper analyzes the performance of the photovoltaic (PV) and thermoelectric generator (TEG) systems for the same reason. The prototype system includes PV, TEG, a heat sink, and a Hofmann Voltameter. The output of the PV–TEG is initially connected to a Hofmann Voltmeter to conduct electrolysis of groundwater and seawater. Further, a heat sink is pasted on each TEG using a thermally conductive adhesive. Then again, PV–TEG/heat sink output is connected to the Voltameter for electrolysis on the same water source samples. Finally, the comparative exploration is done to give insights into PV–TEG applications for hydrogen production. The explorative study shows a DC voltage gain of up to 1.77 times concerning TEG alone. Further, at 849 W/m² of solar irradiance, it is observed that hydrogen production varies from 0.1552 to 0.1639 bar/sec in the case of groundwater and that of seawater goes from 0.1362 to 0.1413 bar/sec when PV–TEG with a heat sink is connected to Voltameter. Graphic abstract
... When there is a temperature difference between the two ends of the thermoelectric plate, thermal energy can be directly converted into electrical energy and output. As a green power generation method, TEG has the advantages of simple structure, durability, no vibration, no noise, no pollution, and high reliability [2]. In recent years, with the development of new high-performance thermoelectric materials and reliable thermoelectric generators [3], the level of research on TEG technology has gradually improved. ...
Article
Full-text available
One of the significant factors contributing to high energy consumption is the unutilized waste heat from flue gas in industrial boilers. Thermoelectric generator (TEG) technology can directly convert thermal energy into electrical energy, and has been gradually applied in the field of waste heat recovery due to its simple and reliable structure, environmental protection, and other advantages. In this paper, a thermoelectric generator simulation system of boiler flue gas waste heat is proposed. The experimental platform is designed by simulating the flue gas waste heat temperature condition of boiler, and the structure of cold end module and hot end module is optimized. During the experiment, the fixed temperature difference was set at 120 °C (hot end:150 °C~cold end: 30 °C). An analysis is conducted on the volt-ampere characteristics and output power of the TEG module. The output characteristics of the TEG system are analyzed under the conditions of variable load, constant load, different pump speed, different heat dissipation modes, and series and parallel connection method. The results show that the experimental platform can instantaneously and accurately test the output parameters of the TEG system, and ensure the intended design requirements. When the ratio of the load resistance to the internal resistance of the TEG module is approximately 1–1.15, the output power of the system reaches its maximum. In order to optimize the output power of the TEG system, a power prediction-based adaptive variable step size maximum power point tracking (MPPT) algorithm is introduced. Additionally, a corresponding mathematical model is formulated. Simulations demonstrate that the time of the improved algorithm to reach the stable maximum power point is 1.54 s faster than that of the traditional algorithm. The improved MPPT algorithm satisfies the criteria for speed and accuracy, diminishes superfluous energy waste, and enhances the overall system efficiency. The research results have certain guiding significance for the design and application of subsequent TEG system.
Article
This paper finds that PV-TEG systems are well-researched. Different techniques are applied to regulate the heat (or dissipate the heat) in these systems. However, the research shows insufficient data regarding TEG as a heat-utilizing and heat-to-electrical energy-converting agent. So, this paper shows the usage of graphite sheets and heat sinks with PV-TEG systems and their comparison. The investigation demonstrates that at 823 W/m2 of solar irradiation, PV-TEG without heat dissipators transforms PV backside heat into TEG DC voltage of 0.727 V; PV-TEG with graphite sheet TEG voltage is 0.889 V; and PV-TEG with heat sink TEG voltage is 1.476 V. The PV-TEG with graphite sheet system improves TEG voltage up to 0.241 V and DC voltage gain up to 0.75 times for TEG alone. In addition, the temperature difference between the hot side and cold side is up to 3.92 C. PV-TEG with a heat sink system improves TEG voltage up to 0.749 V, provides a DC voltage gain of up to 1.78 times for TEG alone, and increases the temperature difference between the hot side and the cold side up to 5.96 °C.
Article
Full-text available
This paper proposes a technique to improve the performance of a thermoelectric element (TEE) using an external magnetic field. The influence of the external magnetic field is investigated with respect to the output electrical power and thermal efficiency of the TEE as a generator, while for the TEE as a cooler, the coefficient of performance, heat absorbed, temperature of the cold side, and current input of the TEE are also examined. From analytical and numerical investigations carried out on bismuth telluride TE material, we reveal that the presence of the magnetic field significantly improves the performance of the TEE operating as a generator below the Curie temperature. Furthermore, in the case of the TEE operating as a generator, we observe a gain of about \(75\,\%\) and \(81.25\%\) corresponding to the electrical output power and thermal efficiency, respectively, whereas in the case of a TEE cooler, its coefficient of performance is significantly improved by \(94.73\,\%\).
Article
Full-text available
In this paper, we present and discuss different aerodynamic solutions available for European tractor and semi-trailers configurations along with summarizing their impact on the aerodynamic drag that contributes to the reduction of fuel consumption. Combinations of different aerodynamic solutions have been presented, and conclusions have been drawn from comparative studies and experience of the participating partners of Project TRANSFORMERS, but no CFD simulations have been done for the same solutions. The paper concludes by proposing configurations of aerodynamic measures for various configurations and how they can be cost-effectively adapted to improve the efficiency in road freight logistics networks and reduce emission.
Article
Full-text available
Growing energy demands are driving people to generate power in every possible way. New energy sources are needed to plug the energy gap. There is a growing interest in distributed energy generation due to its remarkable advantages such as flexibility, reliability, adaptability and minimal transmission losses. Thermoelectric generators (TEGs) are one such distributed power source that relies on thermal energy for electricity generation. The current review focusses on the design and optimization of TEGs to maximize the power output from the available thermal sources. The basic principle of thermoelectricity generation and suitable architecture for specific applications are explained with an overview of materials and manufacturing processes. Various cooling techniques to dissipate heat from the cold side and their influence on overall efficiency are reviewed in this work. Applications of TEGs for powering biomedical sensors have been discussed in detail. Recent advancements in TEGs for various implantable devices and their power requirements are evaluated. The exploitation of TEGs to generate power for wearable sensors has been presented, along with published experimental data. It is envisioned that this study will provide profound knowledge on TEG design for specific applications, which will be helpful for future endeavours.
Article
Full-text available
This paper established a three-dimensional numerical model of segmented thermoelectric generator (TEG) for engines waste heat recovery. The thermoelectric materials are low-temperature Bi2Te3 and medium-temperature skutterudite. Two modes of segmented ratios of p-leg and n-leg are analyzed based on different boundary conditions, such as heat source temperature, cold source temperature and ratios of two thermoelectric materials. Suggestions on improving the output power and conversion efficiency are given based on the analysis. Results show that the performance of a segmented TEG with higher ratios of skutterudite is significantly better than that of a traditional TEG under larger temperature gradient. Enhancing the heat transfer from the heat source is a more effective method of improving the output power. Considering different thermal conductivity, the optimal Bi2Te3 ratio of p-leg is a little higher than that of n-leg. In addition, the design of segmented TEG should consider the temperature range of the heat and cold source to improve the output power and conversion efficiency to the utmost extent.
Article
Full-text available
The performance of thermoelectric materials has been improved in the last years by a significant reduction of the thermal conductivity, but this parameter has now reached its (amorphous) limit, which makes critical the enlargement of the power factor (PF). Here we present a new concept, based on a hybrid system formed by a porous thermoelectric solid permeated by a liquid with an inert salt dissolved (electrolyte), which can be designed to significantly enhance the PF. The concept is demonstrated in an Sb-doped SnO2 porous film permeated with different inert salts (X+(BF4)-, X+=Li+, Na+, K+) dissolved in 3-methoxypropionitrile (3-MPN). A 61.9% average decrease in the electrical resistivity without a significant variation of the Seebeck coefficient was found in the case of 1 M LiBF4 in 3-MPN. As a result, the PF was remarkably increased 3.4 times. On the other hand, 1-Butyl-3-methylimidazolium (BMI+Y-, Y-=I-, BF4-) ionic liquids were also employed as electrolytes. Using the BMII ionic liquid, the electrical resistivity showed a more significant average decrease of 82.5%, however, the absolute value of the Seebeck coefficient was reduced by 35%, finally resulting in an average 2.4 times enhancement of the PF. The large enhancements achieved are attributed to the modification of the electrostatic environment of the porous solid by the ions in the electrolyte at the solid-liquid interface. These results establish a new strategy for the significant improvement of the PF which is not restricted to certain materials and can be potentially applied widely.
Article
A thermoelectric generator used to recover energy from the exhaust gases of a heavy-duty freight vehicle is numerically simulated and optimized using two different algorithms: a gradient-based search method, and a global and local direct search method. The model uses a one-dimensional finite volume method to calculate the temperature and the convective heat transfer of the hot and cold fluids along the streamwise direction, as well as the pressure drop, and employs a thermal resistance network and a global energy balance to account for the heat transfer and the thermoelectric effect. Three different heat exchanger configurations are studied with plain, offset strip or triangular fins. The objective function is the net output power, and the geometrical dimensions of the heat exchanger and thermoelectric units are taken as design variables. The results show that the height of the fins, the spacing between them, the height and side length of the cross section of the thermoelectric units, and the distance between the legs are critical parameters in the optimization process. The thickness of the fins, wall ducts, electrical conductor and ceramic strips should be as small as possible, as well as the height of the cooling wall duct. The two optimization methods yield similar optimal solutions.
Article
The consumption of energy continues to increase at an exponential rate, especially in terms of conventional automobiles. Approximately 40% of the applied fuel into a vehicle is lost as waste exhausted to the environment. The desire for improved fuel efficiency by recovering the exhaust waste heat in automobiles has become an important subject. A thermoelectric generator (TEG) has the potential to convert exhaust waste heat into electricity as long as it is improving fuel economy. The remarkable amount of research being conducted on TEGs indicates that this technology will have a bright future in terms of power generation. The current study discusses the optimal design of the automotive exhaust TEG. An experimental study has been conducted to verify the model that used the ideal (standard) equations along with effective material properties. The model is reasonably verified by experimental work, mainly due to the utilization of the effective material properties. Hence, the thermoelectric module that was used in the experiment was optimized by using a developed optimal design theory (dimensionless analysis technique).
Article
Abstract The waste heat of automotive exhaust gas would be directly transferred into electricity by thermoelectric modules (TEM) because of the temperature difference between heat exchanger and water tank. For the vehicle thermoelectric generator (TEG), the electrical power generation was deeply influenced by temperature difference, temperature uniformity and topological structure of TEG. In previous works, increasing the difference of temperature would significantly enhance the power generation of TEG and inserted fins were always applied to enhance heat transfer in heat exchanger. However the fins would result in a large unwanted back pressure which went against to the efficiency of the engine. In current studies, in order to enhance heat transfer rates and to avoid back pressure increase, a heat exchanger containing cylindrical grooves (the depth-to-width ratio is 0.25) on the interior surface of heat exchanger was proposed. The cylindrical grooves could increase the heat transfer area and enhance the turbulence intensity, meanwhile there was no additional inserts in the fluid to block the flow. The surface temperatures of water tank and heat exchanger with three internal structures, such as grooved surface, flat surface and inserted fins, were studied by numerical simulation at each row of thermoelectric modules. The results showed that comparing to other structures, heat exchanger with cylindrical grooves could improve the TEG efficiency at a low back pressure. The influence of the channel height on the TEG performance was investigated and the TEG with a channel height of 8 mm showed the best overall performance. It was also found that a portion of heat exchanger uncovered with thermoelectric modules at the downstream section could make the upstream section hotter and make full use of each TEM.
Article
A comprehensive mathematical model is proposed to calculate the optimal leg length and cross-section area of TEG unit to maximize the peak output power. The model shows that for a TEG unit, there exists an optimal ratio of leg length and leg cross-section area corresponding to the maximum peak output power with a convective thermal boundary condition, and the optimal leg length and cross-section area can be further calculated based on the optimal ratio. The mathematical model is also validated in this paper, and the corresponding error is within a reasonable range. Moreover, the effects of the leg length and leg area on the peak output power, the peak output power density and the efficiency of TEG unit with different thermal boundary conditions are also discussed. This study will provide guidance for the structure design optimization of TEG unit.
Article
Semiconductor thermoelectric generation technology has a promising application for waste heat recovery and is becoming a noticeable research field. This paper presents a mathematical thermoelectric generator (TEG) model applicable to engine exhaust gas heat recovery. This model considers the conduction along the heat exchanger and the contact thermal resistance of the exchanger plate between the fluid channel and the thermoelectric modules, which were neglected in a previous numerical analysis. Considering that there is an optimum module size corresponding to the maximum power output in a TEG system, the effects of heat transfer characteristics between the fluid channel and the thermoelectric modules on the optimal thermoelectric performance are investigated by using an engineering equation solver (EES) program. Numerical results show that there is a larger optimal area and a lower peak power when considering the conduction and contact thermal resistance factors mentioned above compared to traditional results, which are determined by considerably different factors such as internal heat transfer and temperature distributions. Moreover, an optimal thermal conductivity coefficient that corresponds to the peak power output exists when the conduction along the exchanger is considered. In the case when conduction along the exchanger is not considered, higher the thermal conductivity, the better is the thermoelectric performance. A thin-plate exchanger is recommended in the TEG system owing to its high power output. The optimal module area increases linearly, and the maximum power output exhibits an obvious decrease with an increase in the contact thermal resistance.