ArticlePDF Available

Innovative design for thermoelectric power generation: Two-stage thermoelectric generator with variable twist ratio twisted tapes optimizing maximum output

Authors:

Abstract

In recent years, considerable effort has been dedicated to the development of highly efficient thermoelectric generators for waste heat recovery and thermoelectric power generation. In this study, we present employing twisted tapes with variable twist ratio to enhance thermal energy extraction efficiency, coupled with two-stage thermoelectric modules for heat-to-electricity conversion, resulting in a substantial increase in the power output of the thermoelectric generator. We established an experimental system that validated the superior power generation and heat recovery characteristics of the two-stage thermoelectric generator. Building upon these findings, we propose further optimizing the variable twist ratio twisted tapes to enhance the power output. We investigated the impact of tape pitch ratio, twist ratio, and twist ratio variation range on thermoelectric performance. Experimental results indicate that the influence of flow instability is more pronounced than that of swirl intensity, and twist tapes with the maximum twist ratio variation rate yield the highest net output power. Compared to an unmodified thermoelectric generator, the two-stage thermoelectric generator employing twist tapes with a twist ratio increase from π to 3π achieves a maximum net output power gain of up to 100%. These findings provide a practical framework for integrating innovative power generation modules and optimized heat exchanger designs into the application of waste heat recovery thermoelectric generators, marking a significant advancement in the field of thermoelectric generators.
Applied Energy 363 (2024) 123047
0306-2619/© 2024 Elsevier Ltd. All rights reserved.
Innovative design for thermoelectric power generation: Two-stage
thermoelectric generator with variable twist ratio twisted tapes optimizing
maximum output
Wenlong Yang
a
,
b
, Chenchen Jin
b
, Wenchao Zhu
a
,
c
, Changjun Xie
a
,
b
,
*
, Liang Huang
b
, Yang Li
d
,
Binyu Xiong
b
a
Hubei Key Laboratory of Advanced Technology for Automotive Components, Wuhan University of Technology, Wuhan 430070, China
b
School of Automation, Wuhan University of Technology, Wuhan 430070, China
c
State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
d
Department of Electrical Engineering, Chalmers University of Technology, Gothenburg 41258, Sweden
HIGHLIGHTS
Two-stage thermoelectric modules boosts power output.
Variable twist ratio twisted tapes to boost thermal energy extraction efciency.
Impact of tape pitch ratio and variable twist ratio on thermoelectric performance.
Optimized thermoelectric generator net output power doubled.
ARTICLE INFO
Keywords:
Two-stage thermoelectric generator
Variable twist ratio
Twisted tape
Waste heat recovery
Heat transfer enhancement
ABSTRACT
In recent years, considerable effort has been dedicated to the development of highly efcient thermoelectric
generators for waste heat recovery and thermoelectric power generation. In this study, we present employing
twisted tapes with variable twist ratio to enhance thermal energy extraction efciency, coupled with two-stage
thermoelectric modules for heat-to-electricity conversion, resulting in a substantial increase in the power output
of the thermoelectric generator. We established an experimental system that validated the superior power
generation and heat recovery characteristics of the two-stage thermoelectric generator. Building upon these
ndings, we propose further optimizing the variable twist ratio twisted tapes to enhance the power output. We
investigated the impact of tape pitch ratio, twist ratio, and twist ratio variation range on thermoelectric per-
formance. Experimental results indicate that the inuence of ow instability is more pronounced than that of
swirl intensity, and twist tapes with the maximum twist ratio variation rate yield the highest net output power.
Compared to an unmodied thermoelectric generator, the two-stage thermoelectric generator employing twist
tapes with a twist ratio increase from
π
to 3
π
achieves a maximum net output power gain of up to 100%. These
ndings provide a practical framework for integrating innovative power generation modules and optimized heat
exchanger designs into the application of waste heat recovery thermoelectric generators, marking a signicant
advancement in the eld of thermoelectric generators.
1. Introduction
Various energy-intensive processes, such as industrial activities and
power generation, often result in the wasteful dissipation of excess heat,
leading to a signicant squandering of energy and resources. Waste heat
recovery, a pivotal research domain garnering global attention, presents
an avenue for enterprises to diminish energy consumption, carbon
emissions, and costs. The conversion of waste heat into efcient elec-
trical power is considered among the most optimal techniques for waste
heat utilization, given the substantial volume of waste heat, where even
marginal advancements may yield profound changes [1].
* Corresponding author at: Hubei Key Laboratory of Advanced Technology for Automotive Components, Wuhan University of Technology, Wuhan 430070, China.
E-mail addresses: wenlongyang@whut.edu.cn (W. Yang), jackxie@whut.edu.cn (C. Xie).
Contents lists available at ScienceDirect
Applied Energy
journal homepage: www.elsevier.com/locate/apenergy
https://doi.org/10.1016/j.apenergy.2024.123047
Received 30 December 2023; Received in revised form 23 February 2024; Accepted 15 March 2024
Applied Energy 363 (2024) 123047
2
Thermoelectric Generators (TEGs) emerges as a promising green
waste heat recovery technology, directly transforming thermal energy
into electricity. Thermoelectric generators possess unique advantages
such as solid-state conversion, absolute noiselessness, low maintenance
costs, and structural simplicity [2]. The solid-state conversion is ach-
ieved through Thermoelectric Modules (TEMs), utilizing the Seebeck
effect to directly convert temperature differentials into electrical energy.
Research on waste heat TEGs has been extensively documented, pri-
marily concentrated in the domains of automotive and marine engines
[3].
Despite the increasing prevalence of electric vehicles [4], fossil fuel-
powered automobiles emit substantial waste heat through their exhaust
systems. Due to the limitations of Carnot cycle efciency, this results in
energy wastage and environmental issues. Even a modest proportion,
such as 6%, converted into electricity could reduce fuel consumption by
10% [5], offering numerous benets to both the environment and en-
ergy efciency. Hence, efcient recovery of vehicle waste heat is
imperative to reduce emissions and enhance energy efciency. Liu et al.
[6] installed four TEGs on an off-road vehicle, achieving a maximum
output power of 944 W through road tests and dynamometer experi-
ments. Zhang et al. [7] employed a TEG constructed with nano-
structured materials for recovering waste heat from automotive diesel
engines, achieving 1 kW of electricity production with a thermoelectric
efciency of 2.1%. These studies underscore the feasibility and vast
application prospects of thermoelectric generators for vehicle waste heat
recovery. However, the generation of such signicant electrical power
comes at the cost of exacerbated weight and economic considerations
[8]. Substantial improvements in power density, economic viability, and
conversion efciency are necessary to effectively promote the wide-
spread application of automotive thermoelectric generators.
Researchers are diligently enhancing the performance of TEGs from
multiple perspectives, encompassing the development of novel ther-
moelectric materials [9], innovative thermoelectric module structures
[10], optimization of circuit topologies [11], augmented heat transfer at
the hot end [12], and optimization of heat dissipation [13]. Through
these investigations, fundamental parameters governing the perfor-
mance of exhaust waste heat recovery TEGs have gradually come to
light. Constrained by the low conversion efciency of current commer-
cial thermoelectric materials (below 5%), achieving 20 W of electricity
from a single TEM necessitates the passage of at least 400 W of heat at
the hot end [14]. This is attributed to the fact that, aside from the
intrinsic thermoelectric conversion capability of the material itself, the
power generation is contingent upon the temperature differential at
both ends. However, due to the remarkably limited convective heat
transfer coefcient at the gas-solid interface, collecting a heat ow
exceeding 400 W for each TEM proves to be particularly challenging [1].
Consequently, the heat exchange between the gas and the thermoelec-
tric module emerges as a pivotal factor determining the performance of
thermoelectric generators for exhaust waste heat recovery, constituting
one of the foremost challenges in the current landscape [15].
Through extensive research, researchers have gradually identied
solutions to the issue. A commonly employed method is optimizing the
arrangement of internal heat exchangers, aiming to minimize the ther-
mal resistance of the exchangers. The desired outcome is the reduction
of the temperature differential between the hot gas and the thermo-
electric module surface, ultimately achieving the goal of enhancing the
heat extraction rate and power generation. One of the simplest and most
effective approaches to optimize the heat exchanger arrangement in-
volves the insertion of ow disturbance elements [16]. He et al. [17]
investigated the enhanced heat transfer effects of plate n heat ex-
changers in the context of waste heat recovery TEGs from heavy-duty
diesel engines. The results indicated that the spacing and height of the
ns had a more signicant impact than the thickness of the ns,
although the study did not account for the inuence of ns on exhaust
pressure drop and net power. Pujol et al. [18] explored the thermo-
electric and ow resistance characteristics of a single TEG under forced
convection conditions, employing plate ns. They found that the net
output power was less sensitive to changes in n thickness than changes
in n spacing. Chen et al. [19] conducted a comparative study on the
effects of pin ns and plate ns in enhancing TEG performance. They
discovered that both pin ns and plate ns, compared to smooth pipes
without ns, could multiply the output power of the thermoelectric
module. Following numerical optimization, the output power of pin ns
could be up to 24.14% higher than that of plate ns. Zhao et al. [20]
investigated the impact of inserting perforated bafes into the heat
exchanger on thermoelectric performance. They found that the optimal
installation position could increase the output power by 73.4%, with
exhaust temperature and ow rate having minimal inuence on the
optimal position.
Furthermore, lling the heat exchanger with foam metal not only
maximizes the heat transfer coefcient but also improves the tempera-
ture uniformity along the TEG, alleviating the power mismatch issue to a
Nomenclature
c
p
specic heat capacity, W/(m
2
K)
H pitch length, m
I current, A
L tape distance, m
m mass ow rate, g/s
P power, W
Q
h
hot end heat ow, W
R
L
load resistance, Ω
T temperature, K
U voltage, V
V volume ow rate, m
3
/h
W tape width, m
Δp pressure drop, Pa
ΔT temperature difference, K
Greek symbols
η
efciency, %
ρ
density, g/m
3
σ
maximum error
Subscript
out output value
p pumping value
net net value
ex exhaust gas
te thermoelectric element
hr heat recovery
in inlet value
out outlet value
max maximum value
Abbreviations
TEG thermoelectric generator
TEM thermoelectric module
ITR increasing twist ratio
DTR decreasing twist ratio
TR twist ratio
PR tape pitch ratio
TT twisted tape
W. Yang et al.
Applied Energy 363 (2024) 123047
3
certain extent [21]. Li et al. [22] analyzed the performance of a ther-
moelectric generator with the insertion of porous foam copper in the
central ow region of the heat exchanger. Compared to a thermoelectric
generator without foam metal insertion, the output power of the TEG
using high-porosity, high-volume-ratio porous foam copper could be
increased by up to 2.3 times. However, the pressure drop would rise to
over 10 kPa, posing signicant negative impacts on the engine.
In addition to inserting various structures of ow disturbance devices
in the ducts, enhancing the heat transfer between the gas and the ther-
moelectric module can also be achieved by improving the shape of the
TEGs channels [23]. Luo et al. [24] proposed a convergent heat
exchanger where the hot-side channels converge along the direction of
uid ow. The output power of the convergent TEG was 5.9% higher
than that of a traditional at-plate TEG. Addressing a circular heat
source, Yang et al. [25] introduced a concentric circular thermoelectric
generator, where the center of the hot-side channel was lled with a
cylinder, reducing the uid space and thereby increasing the heat
transfer coefcient. Results indicated that, compared to a smooth heat
exchanger, the net output power of the concentric circular TEG could
increase by up to 65%. The commonality in these studies lies in reducing
the cross-sectional area of uid channels or increasing uid turbulence
to enhance the heat transfer coefcient. However, the improvement in
heat transfer is often accompanied by deterioration in other aspects
[26], such as friction coefcients. Currently, there is a lack of a widely
accepted heat transfer enhancement method that can effectively in-
crease the heat transfer coefcient while maintaining low pressure drop.
Net power is typically used to represent the coupled effects of heat
transfer enhancement in the TEG and the additional pressure drop. The
extra pressure drop results in pumping power losses for the engine. Net
power is generally the difference between TEG output power and
pumping power.
Due to its excellent performance, light weight, simplicity of instal-
lation, and low manufacturing/operating costs, the insertion of twisted
tapes (TTs) has recently become one of the most popular heat transfer
enhancement devices [27,28]. Twisted tapes are typically made of
metal, twisted into specic shapes and sizes, and inserted into the ow
channels. They are also considered vortex generators, serving as tur-
bulence inducers to convey vortex ow, resulting in an increase in the
heat transfer coefcient. There have been few reports on the application
of twisted tapes in TEGs.
Zhu et al. [29] inserted a typical TT into the circular tube heat
exchanger of a annular TEG and optimized the key parameters of the
tape using a genetic algorithm. The results showed that, compared to a
smooth pipe, under the optimal tape structure, the net power and ef-
ciency increased by 10.41% and 22.51%, respectively. They installed a
twisted tape in the heat exchanger that were much smaller in diameter
than the pipe, hence the improvement in TEG performance was not very
pronounced. Researchers have conducted studies on heat transfer
enhancement by inserting TTs in both circular and square tubes [30].
The results indicate that using TTs is an economical heat transfer
enhancement solution. Moreover, due to the high aspect ratio of rect-
angular ducts, their heat transfer increment is higher than that of cir-
cular ducts. This suggests that twisted tapes in at-plate heat exchangers
can achieve better performance. Karana et al. [31] investigated the ef-
fects of different pitch ratios, helix angles, and twist ratios on the ther-
moelectric performance of a diesel engine TEG. The results showed that
the optimal conguration was a pitch ratio of 8, a helix angle of 60, and
a twist ratio of 4, yielding the maximum net output power. At an engine
speed of 1700 rpm and a torque of 60 Nm, the net power could be
increased by 82% compared to a smooth heat exchanger. There is still
further potential for improvement, as there have been extensive reports
indicating that modifying the twisted tape can lead to better heat
transfer performance [28].
In addition to enhancing heat transfer, improving the thermoelectric
modules is an effective method to achieve higher power output within
the limited volume constraints of a TEG [32]. To increase power density,
a promising solution is the implementation of two-stage TEGs [33]. This
conguration involves stacking two typical thermoelectric modules
together, electrically connected in series or parallel, with typically only
a ceramic layer serving as an insulator between them. Recently, research
on two-stage thermoelectric modules has garnered increasing interest
[34,35], as it allows for better electrical energy output within the same
footprint. However, much of the research has been conned to numer-
ical simulations and simulation analyses. In the context of exhaust waste
heat recovery TEGs for internal combustion engines, there is a lack of
experimental evaluation and validation of the thermoelectric properties
of two-stage thermoelectric modules.
Table 1 summarizes recent experimental studies on TEGs used for
internal combustion engine exhaust waste heat recovery. The table in-
cludes information on the thermoelectric module sizes, quantities, heat
transfer enhancement measures, and the obtained output performance.
From the table, it is evident that all studies discussed improvements in
thermoelectric performance using single-stage thermoelectric modules.
However, detailed experimental data on the performance of two-stage
thermoelectric generators and twisted tape-enhanced heat transfer are
still lacking, making reliable optimization studies challenging through
numerical simulations.
This study differs from previous research in two aspects. Firstly, a
two-stage thermoelectric module was manufactured and implemented
to recover exhaust heat energy. Compared to typical thermoelectric
modules, it allows for higher power output within the same footprint.
Secondly, we inserted our proposed twisted tapes with variable-twist
ratio in the heat exchanger to enhance heat transfer. This swirl-
inducing device achieves better heat transfer enhancement. The study
explores the impact of twisted tape congurations on the output power,
conversion efciency, pressure drop, net power, and heat recovery rate
of a two-stage thermoelectric generator under different engine condi-
tions. The results of this research will provide valuable guidance for the
design and optimization of two-stage thermoelectric generators and
twisted tape heat exchangers.
2. Experimental platform development
2.1. Conguration of two-stage TEG
The schematic representation of the TEG is illustrated in Fig. 1(a) and
(b). It comprises a plate-type heat exchanger, two thermoelectric mod-
ule layers, and four radiators. The channel thickness of the plate-type
heat exchanger is 1.5 mm, with dimensions of 320 mm (length) ×
120 mm (width) ×20 mm (height). Twenty-four two-stage thermo-
electric modules are arranged in a 2 ×6 conguration at both ends of the
heat exchanger, interconnected in series. Surrounding the outer surface
of the thermoelectric modules are four radiators, each measuring 300
mm (length) ×50 mm (width) ×12 mm (height), covering every six
TEMs with a cooling conduit. Apart from the TEMs, other principal
components are constructed from aluminum alloy.
Two-stage thermoelectric modules were fabricated based on typical
commercially available Bi
2
Te
3
-based TEM, as illustrated in Fig. 1(c) and
(d). Under the condition of a hot-end temperature of 523 K, the pa-
rameters of a standard commercial single-stage thermoelectric module,
provided by the supplier, are shown in Table 2. The external dimensions
of the two-stage thermoelectric module are 50 mm (length) ×50 mm
(width) ×6.8 mm (height). This two-stage thermoelectric module con-
sists of upper and lower parts, with each level comprised of 198 p-type or
n-type thermoelectric legs (each measuring 1.4 mm in length, 1.4 mm in
width, and 1.6 mm in height) connected in series between 0.3 mm thick
copper electrodes. In total, there are 396 semiconductor legs, forming
198 pairs of pn thermocouples. The outer surface of the two-stage
module and the intermediate region between the two stages are insu-
lated by 0.8 mm thick Al
2
O
3
insulating plates. The Seebeck coefcient
and resistivity are approximately ±190230
μ
V/K and 0.85 mΩ⋅cm,
respectively. Considering the thermal properties of the solder used for
W. Yang et al.
Applied Energy 363 (2024) 123047
4
connecting the thermoelectric legs to the copper electrodes and the
optimal ZT temperature range of BiTe materials, the maximum allow-
able surface temperature is 523 K.
2.2. Twisted tape with variable twist ratio
The installation of twisted tapes is exceedingly straightforward,
accomplished through the insertion into the pipeline. To facilitate the
insertion of the helical tapes, the width (W) of the tapes must align with
the height of the pipeline [28], ensuring a secure fastening within the
heat exchanger channels. When replacement becomes necessary due to
maintenance, structural upgrades, or other reasons, the inserted tapes
can be easily retracted. Following modication, the same installation
method, as described earlier, can be maintained. In this investigation, all
tested twisted tape inserts are crafted from aluminum, possessing a
width of 20 mm, a length of 300 mm, and a thickness of 1 mm.
The pitch length and twist ratio directly inuence the performance of
the twisted tape. The pitch length (depicted in Fig. 2) typically refers to
the length (H) of the tape that undergoes a 180rotation, used in
calculating the twist ratio. The Twist ratio (TR) determines the twisting
frequency of the tape and is calculated as the ratio of pitch length to tape
radius (TR =H/R). Furthermore, within the internal structure of a at-
type heat exchanger, multiple tapes can be inserted since multiple-twist
tapes yield higher heat transfer rates compared to single-twist tapes
[42]. The distance between adjacent tapes, denoted as L, denes the
density of the twisted tape inserts within the nite heat exchanger
channel, represented by the tape pitch ratio (PR =L/W). Fig. 2 illus-
trates the key parameters of the twisted tapes studied in this research.
Twisted tapes have the capacity to generate heightened heat transfer
rates, albeit at the cost of increased frictional losses. To strike a more
Table 1
Recent experimental studies on exhaust heat recovery TEGs for internal combustion engines.
TEM
structure
TEM
quantity
TEM size (length ×width
×height)
Methods to enhance heat transfer Maximum output
power
Maximum
efciency
Pressure
drop
Ref.
Single-stage 24 40 ×40 ×3.3 mm
3
Staggered pin ns 96.6 W 3.95% 0.37 kPa [14]
Single-stage 8 55×51.5 ×3.3 mm
3
Partially lled with copper foam 36 W N/A 5.4 kPa [22]
Single-stage 20 56×56 ×5.5 mm
3
Typical twisted ribs 94.7 W N/A N/A [31]
Single-stage 40 44 ×44 ×3.6 mm
3
Stainless steel perforated plates and plate ns 119 W 2.8% 1.4 kPa [36]
Single-stage 10 55×55 ×4 mm
3
Rectangular winglet longitudinal vortex
generators 7.1 W 1.1% N/A [37]
Single-stage 40 50 ×50 ×3.8 mm
3
Smooth at heat exchanger (4 mm in height) 29.5 W 1.25% 0.12 kPa [38]
Single-stage 30 60 ×60 ×3.3 mm
3
Flow straightener and plate ns 83 W 2.5% 1.9 kPa [39]
Single-stage 18 40 ×40 ×3.6 mm
3
Conical heat exchanger structure with
perforated plate and n 98.8 W 2.6% 2.05 kPa [40]
Single-stage 32 40 ×40 ×3.3 mm
3
Porous copper foam 5.8 W 2.05% 0.81 kPa [41]
Fig. 1. (a) TEG structure diagram. (b) TEG structure splitting diagram. (c) Two-dimensional diagrams of single-stage and two-stage TEM. (d) Manufactured two-
stage TEM.
Table 2
Parameters of single-stage TEM (TEG-19913).
Parameters Cold end temperature
300 K 323 K
Open-circuit voltage (V) 13.0 11.8
Dynamic internal resistance (Ω) 2.8 2.9
Maximum output power (W) 14.5 11.5
Conversion efciency (%) 5.7 5.0
AC internal resistance (Ω, 297 K) 1.58 1.58
Fig. 2. Key parameters of twisted tapes.
W. Yang et al.
Applied Energy 363 (2024) 123047
5
favorable balance between enhanced heat transfer rates and augmented
pressure losses, renement of the twisted tape design is imperative. In
the case of exhaust gas waste heat recovery systems where a signicant
temperature gradient is prevalent along the ow direction of the chan-
nels, the distinct temperature gradients induce uneven potential distri-
bution across the series-connected TEMs, resulting in parasitic power
losses [43]. Therefore, the objective of this study is to ameliorate this
scenario and seek the most efcient heat transfer enhancement measures
while maintaining a lower pressure drop. A novel twisted tape structure
is proposed, characterized by a variable pitch along the length of the
tape. In three-dimensional space, the structure of the twisted tape can be
represented by the following expression:
x=Rcos(2
π
WTR t)
y=Rsin(2
π
WTR t)
z=t
(1)
where t =[0,300].
To construct a twisted tape with variable twist ratio, one simply
needs to linearly increase or decrease 2
π
/(WTR). When the TR gradu-
ally decreases along the uid ow direction, it is termed as a decreasing
twist ratio twisted tape (DTR-TT); conversely, when the TR gradually
increases along the uid ow direction, it is designated as an increasing
twist ratio twisted tape (ITR-TT).
For a standard constant twist ratio twisted tape, a general trend
suggests that as the pitch length decreases, thermodynamic performance
increases. This relationship exhibits an optimal point where the increase
in frictional losses and the improvement in convective heat transfer
reach an optimal balance. Traditionally, researchers often adopt 6as
the optimum TR value [30]. In the realm of variable twist ratio twisted
tapes, it can be perceived as a compromise value, wherein variations in
pitch are explored based on TR =6, investigating the performance of
variable twist ratio twisted tapes under different combinations. Ac-
cording to the expression for variable twist ratio tapes, Eq. (1), its TR
alteration is associated with
π
. Consequently, we have devised specic
specications of variable twist ratio twisted tapes based on
π
, 2
π
, and 3
π
as TR boundaries. Fig. 3 illustrates schematic diagrams of typical con-
stant twist ratio twisted tapes and variable twist ratio twisted tapes.
To handle the structural characteristics of variable twist ratio twisted
tapes in a dimensionless manner, we dene the variable twist ratio
parameter (VTR) as the ratio of the twist ratio at the beginning to that at
the end of the tape. Thus, for three types of increasing twist ratio twisted
tapes, VTR =1/3, 1/2, and 2/3, while for three types of decreasing twist
ratio twisted tapes, VTR =3/2, 2, and 3. This denition can be extended
to variable twist ratio twisted tapes with different structural parameters,
where VTR <1 signies increasing twist ratio twisted tapes, VTR =1
denotes conventional constant twist ratio twisted tapes, and VTR >1
indicates decreasing twist ratio twisted tapes. As VTR approaches 1, the
rate of twist variation diminishes, whereas it increases as VTR deviates
further from 1.
2.3. Apparatus and experimental settings
The fabricated two-stage TEMs were integrated with the variable
twist ratio twisted tapes into the TEG experimental system for the
assessment of thermoelectric performance. By designing the physical
structure of the variable twist ratio twisted tapes in three-dimensional
software and subsequently employing metal 3D printing technology,
we fabricated the DTR-TT and ITR-TT. The experimental system consists
of an air heating system, TEG, a cooling water circulation system, an
electronic load, and a data acquisition system (Fig. 4).
The air heating system utilizes an industrial hot air blower (RY-P-15
A-075) as a simulated heat source, replacing an automobile engine, to
steadily provide the required ow and temperature of high-temperature
gas [24]. The industrial hot air blower comprises a blower, a heater, and
a control circuit. With a maximum power of 15 kW, the industrial hot air
blower draws in ambient air and heats it. The exhaust temperature can
be adjusted within the range of 300633 K with a PID controller (ac-
curacy ±0.5 K). The outlet ow rate of the hot air blower can be regu-
lated within the range of 0240 m
3
/h by adjusting the control panel.
Section 2.1 has provided a detailed description of the customized
TEG, with a detachable side of the hot-end heat exchanger, allowing the
Fig. 3. (a)Variations of the twist ratio along the length and schematic diagrams of (b) constant twist ratio, (c) decreasing twist ratio and (d) increasing twist ratio of
the twisted tapes.
W. Yang et al.
Applied Energy 363 (2024) 123047
6
insertion of the twisted tape. The TEG is connected to the industrial hot
air blower via anges. To minimize thermal resistance, the contact
surfaces between the heat exchanger, thermoelectric modules, and ra-
diators are coated with thermally conductive silicone grease. Bolts are
applied on both sides of the TEG for secure fastening.
The cooling water circulation system primarily consists of a water
tank, a water pump, and pipelines. A chilling unit (SHZ-95B) pumps cold
water from a 50 L water tank through the cold-side heat exchanger of the
TEG, maintaining a low temperature on the cold side of the TEG. The
output ow rate of the chilling unit is approximately 10 L/min, con-
nected to four radiators in the TEG system via rubber hoses. Due to the
large capacity of the cooling water tank, the cooling water temperature
remains nearly constant during the experimental process, achieving
stable output shortly after starting the TEG.
The electronic load and data acquisition system include a vortex
owmeter, a DC power supply unit, an electronic load, and a data
acquisition instrument. To measure the actual exhaust outlet ow rate,
temperature, and heat exchanger outlet pressure, a vortex owmeter
(HW-LUGA14, full range 50480 m
3
/h, accuracy ±1%) is installed at
the outlet of the heat exchanger. A 24 V DC power supply (RIGOL
DP832) powers the vortex owmeter. An electronic load (Array 3721 A,
resolution 0.1 mΩ, accuracy ±0.5%) is connected to the thermoelectric
modules to simulate external load conditions. All experimental data,
including the TEGs output current, temperature, and pressure, are
measured and collected using a data acquisition device (Agilent DSO-X
2024 A).
The hot air blower, TEG, and vortex owmeter are connected using
stainless steel pipes with anges. The hot uid output from the hot air
blower ows through the TEG and vortex owmeter, discharging into
the environment. Furthermore, during the experimental process, the
TEG is enveloped in silicone felt with an extremely low thermal con-
ductivity to minimize heat losses. Given the direct connection between
the hot air blower and the TEG, the temperature at the outlet of the
blower is considered as the inlet temperature of the TEG. This temper-
ature measurement is facilitated by sensors and controllers integrated
within the blower. The outlet temperature is directly measured via the
vortex owmeter. The ow velocities at the TEGs inlet and outlet are
directly measured by the hot air blower and the vortex owmeter,
respectively. The inlet pressure of the TEG is estimated by referencing
the performance curve provided by the manufacturer of the hot air
blower. This curve delineates the relationship between outlet pressure,
airow velocity, and power under various operating conditions. The
outlet pressure of the TEG is directly measured by the vortex owmeter.
2.4. Experimental procedure
The power output and back pressure of the TEG are signicantly
inuenced by the operational state of the engine. Kim et al. [36] con-
ducted tests on a turbocharged six-cylinder diesel engine, recording
operational parameters of the TEG system under various engine modes.
The engine had a cylinder bore of 103 mm, a stroke of 118 mm, a
compression ratio of 17:1, and a maximum power of 110 kW at 2500
rpm. Fig. 5 illustrates the impact of engine load on TEG inlet tempera-
ture and mass ow rate at different engine speeds. In the current study,
to simulate the actual operating conditions of an automotive waste heat
recovery TEG system, six typical vehicle operating conditions from
Ref. [36] were considered, as shown in Table 3. During the experiments,
the outlet temperature and ow rate of the industrial hot air blower
were adjusted through the control panel to operate the TEG under
specic engine exhaust conditions. Furthermore, the inuence of vehicle
speed on the entrance parameters of the cooling system is negligible
[44].
In each run, the TEG initially operated in an open circuit. Firstly, the
cooling water unit was activated to ensure the circulation of cooling
water. Subsequently, the industrial hot air blower was turned on and
adjusted to a xed exhaust condition. After a period of operation, the
outlet temperature of the TEG was dynamically measured, considering it
in a steady state when the rate of change over ve consecutive minutes
was within ±0.2 K. Finally, the electronic load was connected to the
TEM, and parameters such as TEG voltage, current, outlet pressure, and
ow rate were measured and recorded.
2.5. Data analysis
In the experiment, as the industrial hot air blower is directly con-
nected to the TEG, the outlet temperature of the blower is considered as
Fig. 4. Experimental conguration.
Fig. 5. Variations in exhaust temperature and mass ow rate with engine speed
and load.
W. Yang et al.
Applied Energy 363 (2024) 123047
7
the inlet temperature of the TEG (T
ex,in
). The outlet temperature (T
ex,out
)
and volumetric ow rate (V
ex
) of the TEG were measured using the
vortex owmeter. Therefore, the thermal energy extraction rate (Q
h
)
from the exhaust ow in the TEG region is calculated by the following
Eq. [45]:
Qh=mexcpex ΔTex =Vexcpex ΔTex/
ρ
ex (2)
The maximum possible energy extraction rate can be obtained by
using the atmospheric temperature as the reference temperature, as
shown below:
Qh,max =Vexcpex ΔTex,a/
ρ
ex (3)
η
hr =Qh/Qh,max (4)
where ΔT
ex
represents the temperature difference of the exhaust at the
TEG inlet and outlet, and ΔT
ex,a
represents the difference between the
exhaust temperature at the TEG inlet and the ambient temperature. The
ratio of the actual energy extraction rate (Q
h
) to the maximum possible
energy extraction rate (Q
h,max
) indicates the thermal energy recovery
efciency
η
hr
of the TEG.
The output power (P
out
) is dependent on the load resistance (R
L
) and
can be calculated by measuring the current (I) and voltage (U) across the
load resistor. The calculation formula is as follows:
Pout =UI (5)
The thermoelectric conversion efciency
η
te
is expressed as:
η
te =Pout/Qh(6)
When applying a disrupter to the hot-end heat exchanger of the TEG,
to calculate the back pressure power loss, according to [46], it can be
computed as:
Pp=VexΔp=mex Δp/
ρ
ex (7)
where Δp is the pressure drop of the hot uid through the heat
exchanger, and
ρ
ex
is the density of the exhaust gas.
Therefore, the net output power (P
net
) of the TEG is calculated as:
Pnet =Pout Pp(8)
In addition, the measurement error is calculated based on the de-
nition of heat transfer rate and the accuracy of each measurement sensor
[38]. The specic equation is as follows:
Qh=
(
Qh
Vex
σ
Vex)2
+(
Qh
Tex,in
σ
Tex,in)2
+(
Qh
Tex,out
σ
Tex,out)2
(9)
where
σ
represents the maximum error of each measurement parameter.
The calculated Q
h
error is within ±3.17%. According to the above
equation, the uncertainties of the output power, conversion efciency,
and pressure drop are calculated to be ±0.171%, ±3.181%, and ±
0.862%, respectively.
3. Results and discussion
3.1. Output performance of two-stage thermoelectric generators
Primarily, an examination was conducted within the smooth channel
devoid of twisted tapes, exploring the electrical characteristics of single-
stage TEG and two-stage TEG under varying external resistances. To
quantitatively analyze power output, power-load resistance curves for
the TEG system were obtained by altering external loads under diverse
engine modes, as depicted in Fig. 6. Irrespective of whether it was a
single-stage TEG or a two-stage TEG, the power output increased with
escalating engine speed and load. Under engine mode F, the two-stage
TEG outperformed the single-stage TEG, exhibiting a power output
enhancement of 17.3%. The enhancement rate decreases with the in-
crease of exhaust temperature and mass ow rate. This phenomenon
arises because, despite the heightened temperature gradient between
the cold and hot ends of the TEG, the enhancement in the operational
temperature difference for both the upper-level and lower-level TEMs is
not pronounced. For the single-stage TEG, characterized by lower in-
ternal resistance, the increase in temperature gradient yields a more
conspicuous improvement in output power.
With the escalation of external resistance, the output power initially
rises before diminishing. There exists an optimal load resistance,
enabling the TEG to achieve peak output power. Furthermore, from
Fig. 6, it is discernible that the operational mode of the engine signi-
cantly inuences peak output power. The optimal load exhibits a
Table 3
Selected typical TEG operating conditions.
Engine
mode
Engine Load
(MPa)
Speed
(rpm)
Mass ow rate
(g/s)
Temperature
(K)
A 0.4 1000 19.7 470.7
B 0.4 1500 26.3 509.9
C 0.4 2000 41.5 551.1
D 0.6 1000 21.3 533.1
E 0.6 1500 30.4 560.3
F 0.6 2000 49.5 598.8
Fig. 6. Output power vs. load resistance curves of (a) Single-stage TEG and (b)
Two-stage TEG under different engine modes.
W. Yang et al.
Applied Energy 363 (2024) 123047
8
marginal increase with the rise in intake temperature and ow, its
impact on the optimal load is relatively modest. This proves advanta-
geous for two-stage TEG applications, signifying that when the gener-
ator functions as an external power source, frequent adjustments to
external resistance are unnecessary, ensuring that the output power
remains close to its maximum value [47].
The optimum load resistance for the single-stage TEG is 62 Ω, while
for the two-stage TEG, it is 136 Ω. Despite the doubling of thermoelectric
legs in the two-stage TEG compared to the single-stage TEG, the optimal
load resistance is not twice that of the single-stage TEG. This disparity
arises because in the two-stage TEG, the average operating temperature
of the TEMs is lower. This, coupled with the dual effects of the high
resistivity of semiconductor materials at low temperatures and the
mismatch in internal resistance caused by the series connection of 24
two-stage thermoelectric modules, leads to the optimal load resistance
for the two-stage TEG being 2.2 times that of the single-stage TEG.
In theory, the matched load resistance equals the internal resistance
of the TEMs [48]. In contrast to these ideal conditions, this study
measured output power considering mismatches in resistance and
thermal resistance between 24 single-stage and two-stage TEGs, as well
as power losses caused by lead connections and uneven surface tem-
peratures. Consequently, the optimal load resistance does not align
perfectly with the internal resistance provided by the supplier.
Fig. 7(a) contrasts the thermoelectric conversion efciency and heat
recovery efciency of single-stage and two-stage TEGs under different
operating conditions. The results indicate that the two-stage TEG ex-
hibits a slightly higher thermoelectric conversion efciency than the
single-stage TEG, with the maximum conversion efciencies reaching
1.9% and 2.6% for single and two-stage TEG under various conditions.
As the engine speed and load increase, the thermoelectric conversion
efciency also rises, but there is a declining trend in heat recovery
performance. This is due to the fact that the actual energy extraction rate
increases by a magnitude smaller than the maximum possible energy
extraction rate. Both single and two-stage TEGs exhibit similar trends in
heat recovery performance. At the lowest engine speed and load con-
ditions, the heat recovery efciency is maximal, reaching 36.3% and
35% for single and two-stage TEGs, respectively. However, at the
highest engine speed and load conditions, the heat recovery efciency
decreases to 26.3% and 24.6% for single and two-stage TEGs. The heat
recovery performance limits the maximum output power of the TEG.
This can be elucidated by examining the temperature differentials at
the inlet and outlet of the TEG based on the exhaust. Fig. 7(b) delineates
the exhaust gas temperatures at the inlet and outlet of the TEG, as well as
the temperature differential, under various engine operating modes. It is
distinctly observable that, with the augmentation of engine speed and
load, the temperature differential increases, albeit with a discernible
trend of diminishing increments. A comparative analysis of diverse en-
gine conditions reveals that this phenomenon stems from the fact that,
despite the elevated intake temperature and mass ow rate accompa-
nying the escalation of engine speed and load, the rapid ow velocity
results in the exhaust gases bypassing the TEG without undergoing
sufcient convective heat exchange with the TEMs, thereby owing
directly out of the TEG through smooth channel.
3.2. Inuence of tape pitch ratio
The two-stage TEG manifest a notable potential for high power
density in thermal energy recovery [49]. An examination was conducted
on the inuence of employing twisted tape heat exchangers on the
performance of the two-stage TEG. Initially, an investigation was carried
out on the impact of the tape pitch ratio (PR) on the thermoelectric
performance of the two-stage TEG. In the experiments, the PR was set at
3, 2, 1.5, 1.2, and 1, with a twist ratio of
π
. As depicted in Fig. 8(a), the
variation of output power with engine modes under different PR is
illustrated. The insertion of twisted tapes exhibits an augmentation of
thermoelectric performance. In comparison to smooth channels, the
utilization of twisetd tapes at a lling ratio of PR =1, TR =
π
, can
potentially double the output power. Moreover, as the engine operating
conditions intensify, the increase in output power becomes more sub-
stantial. This arises from the elongation of the ow path by the twisted
tapes, prolonging the contact time between the uid and the tube wall,
ensuring their thorough interaction. Furthermore, the twisted tapes
induce secondary ows, segregating and obstructing uid ow, thereby
accelerating the mixing velocity of uid in the near-wall region and
increasing the ow velocity near the wall [27].
As the pitch ratio diminishes, the TEG output power increases. A
smaller tape gap gives rise to multiple swirling ows with greater in-
tensity, leading to more chaotic uid mixing and increased turbulent
kinetic energy. Consequently, a smaller gap provides a thinner thermal
boundary layer. This enhanced momentum transfer increases the heat
extracted from the hot gas, ultimately raising the temperature on the
TEGs hot side, resulting in higher power output.
The disturbances caused by the twisted tapes not only impact heat
transfer but also inuence the friction between the channel surface and
the uid. The increased channel area blocked by the twisted tapes and in
contact with the uid contributes to pressure loss. As shown in Fig. 8(b),
under different PR, the pressure drop varies with engine modes. It is
observable that as the PR decreases, the increase in pressure drop is
more pronounced compared to the increase in power, reaching a
maximum 6.74-fold increase over smooth pipes. This is attributed to the
minimal PR signifying a heat exchanger lled with twisted tapes,
resulting in more complex uid ow and higher ow resistance.
Fig. 8(c) illustrates the impact of PR on the net output power under
Fig. 7. Comparison of (a) Thermoelectric conversion efciency and heat re-
covery efciency and (b) inlet/outlet temperatures and temperature differences
between single-stage TEG and two-stage TEG.
W. Yang et al.
Applied Energy 363 (2024) 123047
9
various engine operating conditions. Net power reects the coupled
inuence of enhanced heat transfer and frictional losses due to the
twisted tapes on TEG performance. It is evident that for different engine
conditions, net power exhibits distinct trends. At lower engine speeds,
net power increases with decreasing PR, albeit with a diminishing rate of
increase. For engine conditions A, B, D, and E, net power is nearly
equivalent at tape pitch ratios of 1.5, 1.2, and 1. As engine speed in-
creases (engine conditions C and F), the substantial increase in exhaust
ow leads to a considerable rise in channel pressure drop, causing net
power to decrease with decreasing PR.
3.3. Inuence of twist ratio
The investigation delves into the impact of constant twist ratio and
variable twist ratio twisted tapes on the output characteristics and
pressure drop of the two-stage TEG system. Fig. 9a illustrates the vari-
ation of output power under different operating conditions for twisted
tapes with PR =1.5, employing constant, decreasing, and increasing
twist ratios. For a uniform pitch, the output power increases with a
decrease in the twist tatio. This is attributed to the enhanced convective
heat transfer with a diminishing TR. Across various operating condi-
tions, the output power with TR =
π
is maximized, exhibiting a 21%
improvement over the twisted tape with TR =3
π
. The smaller TR results
in a greater number of helices, generating stronger swirling ows and
higher turbulence intensity, thereby achieving superior heat transfer. As
the exhaust temperature and ow rate decrease, the enhancement ratio
of output power with TR =
π
compared to TR =3
π
declines from 21% to
16%, owing to the superior thermal performance of twisted tapes at low
Reynolds numbers [30].
For a typical twisted tape, the thermodynamic properties vary with
the degree of looseness of the tape. Larger TR leads to a looser cong-
uration, and vice versa. According to Fig. 9, the DTR-TT of TR =3
π
to
π
is tighter than that of TR =3
π
to 2
π
, so the output power is greater. The
DTR-TT with TR =3
π
to
π
exhibits greater output power than the CTR-
TT with TR =3
π
, yet less than the twisted tape with TR =
π
. However, it
surpasses the output power of the average TR =2
π
twisted tape, indi-
cating the enhanced heat transfer performance of variable twist ratio
TTs compared to constant twist ratio TTs.
Under the same variable twist ratio combination, the output power of
Fig. 8. Inuence of different tape pitch ratios on two-stage TEG performance
under various engine modes: (a) Output power, (b) Pressure drop, and (c)
Net power.
Fig. 9. Output power and pressure drop of constant and variable twist ratios
under different engine modes.
W. Yang et al.
Applied Energy 363 (2024) 123047
10
the ITR-TT is consistently higher than that of DTR-TT. For instance, in
engine mode F, the output power of TR =
π
to 3
π
ITR-TT is 25% higher
than that of TR =3
π
to
π
DTR-TT. This is attributed to the lower pressure
drop upstream in general pipes, resulting in lower pressure loss and
higher temperature gradients. The addition of ITR-TT increases the
number and intensity of swirls upstream, yielding superior heat transfer
performance. Therefore, twisted tapes with shorter twist lengths inser-
ted upstream in the pipe achieve higher TEG output power.
An noteworthy observation is that the output power of TR =
π
to 3
π
ITR-TT is higher than that of TR =
π
to 2
π
ITR-TT. The looser variable
twist ratio TT enables TEG to attain better output performance, in
contrast to the pattern observed with DTR-TT. This is because the non-
uniform distribution of the internal vortex ow eld in TR =
π
to 3
π
ITR-TT (VTR =1/3) is more pronounced than that in TR =
π
to 2
π
ITR-
TT (VTR =1/2). This non-uniform distribution generates a larger
effective driving potential induced by the twisted tape than the one
induced by a uniform tape, resulting in higher heat and momentum
uxes. Hence, for increasing twist ratio twisted tapes, the inuence of
ow instability is more signicant than the inuence of swirl intensity.
Expanding the pitch variation range of variable twist ratio TTs can
enhance heat transfer performance.
Fig. 9(b) illustrates the variation in pressure drop under different
Fig. 10. Theoretical ow patterns under different TR and PR.
W. Yang et al.
Applied Energy 363 (2024) 123047
11
operating conditions for CTR, DTR, and ITR twisted tapes. The trend in
pressure drop mirrors the trend in TEG output power, as the effect of
enhanced heat transfer is accompanied by an increase in friction coef-
cient. The noteworthy difference is that, for DTR-TT, the impact of ow
instability on pressure drop is greater than its impact on heat transfer
performance. The output power and pressure drop of TR =3
π
to
π
DTR-
TT are 2.5% lower and 3.2% higher, respectively, compared to those of
TR =2
π
to
π
DTR-TT. The ow patterns depicted in Fig. 10 aid in un-
derstanding the inuence of twist ratio and tape pitch ratio on uid ow
and heat transfer mechanisms within the heat exchanger channels in this
study [50].
3.4. Net power analysis
The enhanced heat transfer effectiveness of variable twist ratio
twisted tapes surpasses that of typical twisted tapes, but the resulting
additional pressure drop is also signicantly higher. Once the engine
back pressure exceeds a certain level, it not only affects the normal
operation of the engine but also leads to an increase in resistance power
consumption, potentially surpassing the output power generated by the
TEG itself [51]. To assess the overall performance of variable twist ratio
twisted tapes in the TEG system, the impact of different variable twist
ratio TT on the net output power of the TEG was evaluated.
Fig. 11(a), (b), and (c) depict the variation in net output power for
TEGs employing CTR, DTR, and ITR twisted tapes with PR =1.5 under
different operating conditions. For CTR-TT, the net power increases with
a decrease in TR. However, for DTR-TT, the tape with the highest rate of
TR variation results in lower net power, particularly at high engine
speeds. This indicates that the ow instability of DTR-TT leads to infe-
rior overall performance compared to typical tapes, corroborating the
conclusions from Fig. 9. In contrast, increasing twist ratio twisted tape
yields different results. In most operating conditions, TR =
π
to 3
π
ITR-
TT achieves the best thermoelectric performance. In mode D, the net
output power of TR =
π
to 3
π
ITR-TT is 5.9% and 17.1% higher than that
of CTR-TT with twist ratios of
π
and 2
π
, respectively.
Fig. 11(d), (e), and (f) illustrate the variation in net thermoelectric
efciency (
η
net
=P
net
/Q
h
) for TEGs employing CTR, DTR, and ITR
twisted tapes with PR =1.5 under different operating conditions. The
trend in net efciency follows a similar pattern to net power. The net
efciency of ITR-TT is superior to CTR-TT at low engine speeds and
comparable at high engine speeds. From an efciency standpoint, ITR-
TT also represent the most effective enhancement for thermoelectric
performance.
To accentuate the advantages of variable twist ratio twisted tapes,
Fig. 12 compares the ratio of the net power of the TEG system after
enhanced heat transfer to the net power of the TEG system before
Fig. 11. Net power of (a) Constant twist ratio, (b) Decreasing twist ratio, and (c) Increasing twist ratio under different engine modes; as well as net conversion
efciency of (a) Constant twist ratio, (b) Decreasing twist ratio, and (c) Increasing twist ratio under different engine modes.
Fig. 12. Net power ratio of two-stage TEG with twisted tape heat exchanger
and two-stage TEG with smooth heat exchanger.
W. Yang et al.
Applied Energy 363 (2024) 123047
12
enhanced heat transfer [52]. This ratio characterizes the enhancement
in overall TEG performance relative to smooth heat exchanger. As en-
gine speed and load increase, the net power enhancement ratio of
twisted tape decreases. In engine mode A, twisted tapes with twist ratios
of 3
π
, 2
π
,
π
, 3
π
-2
π
, 3
π
-
π
, 2
π
-
π
, 2
π
-3
π
,
π
-3
π
, and
π
-2
π
yield net power
enhancements of 57.7%, 70.5%, 89%, 60%, 72.8%, 78.3%, 64.7%,
100%, and 98%, respectively, compared to a regular smooth heat
exchanger. Except for engine mode F, the enhancement ratio of TR =
π
to
3
π
ITR-TT is superior to that of CTR-TT and DTR-TT in other operating
conditions.
Finally, to highlight the novelty and advancement of this study, we
compare the two-stage TEG with the optimal variable twist ratio twisted
tapes to a typical TEG without heat transfer optimization, as shown in
Fig. 13. The typical TEG employs commercially available single-stage
thermoelectric modules and is equipped with a smooth at-plate heat
exchanger. In the two-stage TEG, the two-stage TEMs developed in
Section 2.1 are utilized, and TR =
π
to 3
π
variable twist ratio twisted
tapes with a tape pitch ratio of 1.5 are inserted into the heat exchanger,
as discussed in Section 2.1.
It is evident that, through the optimization of thermoelectric mod-
ules and heat transfer performance on the hot side, the overall perfor-
mance of the TEG has been signicantly enhanced, despite the increase
in exhaust pressure drop. Although the pressure drop has increased by
1.5 times, the output power, net power, thermoelectric conversion ef-
ciency, heat recovery efciency, and net efciency have all improved
by approximately double. It is worth noting that if the issue of internal
resistance mismatch in the two-stage thermoelectric modules is
addressed, and compatibility with various thermoelectric materials is
achieved for the temperature gradient, the comprehensive performance
of the TEG will see further substantial improvement. This will be pur-
sued in our future research endeavors. As the cost of thermoelectric
modules continues to decrease, there is a potential for the power-to-cost
ratio of two-stage thermoelectric generators to surpass that of single-
stage thermoelectric generators.
4. Conclusions
In this study, we conducted a comprehensive examination of the
experimental system for a exhaust heat recovery thermoelectric gener-
ator, comparing the thermoelectric characteristics of single-stage and
two-stage generators. Additionally, we introduced a method to enhance
heat transfer through the utilization of variable twist ratio twisted tapes
and systematically investigated its impact on the performance of the
two-stage thermoelectric generator. From our research ndings, the
following conclusions can be drawn:
1. In comparison to a single-stage thermoelectric generator, under en-
gine mode F, the two-stage thermoelectric generator leads to a 17.3%
increase in output power, and the optimum load resistance is
enhanced by approximately 2.2 times.
2. The output power and pressure drop of the two-stage thermoelectric
generator both increase with a decrease in tape pitch ratio, with the
pressure drop increase being more pronounced. Compared to a
smooth heat exchanger, the minimum tape pitch ratio results in a
6.74-fold increase in pressure drop. At low engine speeds and loads, a
smaller tape pitch ratio yields higher net output power, while at high
engine speeds and loads, net power decreases with a decrease in tape
pitch ratio.
3. Output power and pressure drop increase with a decrease in twist
ratio. Due to the more signicant inuence of ow instability
compared to swirl intensity, the output power of the twisted tape
with an increased twist ratio from
π
to 3
π
is 25% higher than that of
the twisted tape with a twist ratio increased from
π
to 2
π
.
4. The increasing twist ratio twisted tape with the highest twist ratio
variation rate exhibits the highest net output power. Under engine
operating condition A, the net output power gain can reach a
maximum of 100% compared to the unmodied thermoelectric
generator.
CRediT authorship contribution statement
Wenlong Yang: Conceptualization, Data curation, Writing original
draft. Chenchen Jin: Investigation, Writing original draft. Wenchao
Zhu: Methodology, Software, Validation. Changjun Xie: Funding
acquisition, Project administration, Resources, Writing review &
editing. Liang Huang: Formal analysis, Software, Visualization. Yang
Li: Supervision, Validation, Writing review & editing. Binyu Xiong:
Formal analysis, Methodology, Writing review & editing.
Declaration of competing interest
The authors declare that they have no known competing nancial
interests or personal relationships that could have appeared to inuence
the work reported in this paper.
Data availability
Data will be made available on request.
Acknowledgments
This work is supported by the National Natural Science Foundation
of China (51977164) and the Postdoctoral Fellowship Program of CPSF
(GZC20232011).
References
[1] He J, Li K, Jia L, Zhu Y, Zhang H, Linghu J. Advances in the applications of
thermoelectric generators. Appl Therm Eng 2024;236:121813. https://doi.org/
10.1016/j.applthermaleng.2023.121813.
[2] Burnete NV, Mariasiu F, Depcik C, Barabas I, Moldovanu D. Review of
thermoelectric generation for internal combustion engine waste heat recovery.
Prog Energ Combust 2022;91:101009. https://doi.org/10.1016/j.
pecs.2022.101009.
[3] Saha M, Tregenza O, Twelftree J, Hulston C. A review of thermoelectric generators
for waste heat recovery in marine applications. Sustain Energy Technol Assess
2023;59:101684. https://doi.org/10.1016/j.seta.2023.103394.
Fig. 13. Comparison of thermoelectric performance between single-stage TEG
with smooth heat exchanger and two-stage TEG with ITR-TT heat exchanger
under engine mode E.
W. Yang et al.
Applied Energy 363 (2024) 123047
13
[4] Waseem M, Amir M, Lakshmi GS, Harivardhagini S, Ahmad M. Fuel cell-based
hybrid electric vehicles: an integrated review of current status, key challenges,
recommended policies, and future prospects. Green Energy Intell Transp 2023;2
(6):100121. https://doi.org/10.1016/j.geits.2023.100121.
[5] Ochieng AO, Megahed TF, Ookawara S, Hassan H. Comprehensive review in waste
heat recovery in different thermal energy-consuming processes using
thermoelectric generators for electrical power generation. Process Saf Environ
2022;162:13454. https://doi.org/10.1016/j.psep.2022.03.070.
[6] Liu X, Deng YD, Li Z, Su CQ. Performance analysis of a waste heat recovery
thermoelectric generation system for automotive application. Energ Conver
Manage 2015;90:1217. https://doi.org/10.1016/j.enconman.2014.11.015.
[7] Zhang Y, Cleary M, Wang X, Kempf N, Schoensee L, Yang J, et al. High-temperature
and high-power-density nanostructured thermoelectric generator for automotive
waste heat recovery. Energ Conver Manage 2015;105:94650. https://doi.org/
10.1016/j.enconman.2015.08.051.
[8] Zhu W, Yang W, Yang Y, Li Y, Li H, Shi Y, et al. Economic conguration
optimization of onboard annual thermoelectric generators under multiple
operating conditions. Renew Energy 2022;197:48699. https://doi.org/10.1016/j.
renene.2022.07.124.
[9] Ying P, He R, Mao J, Zhang Q, Reith H, Sui J, et al. Towards tellurium-free
thermoelectric modules for power generation from low-grade heat. Nat Commun
2021;12:1121. https://doi.org/10.1038/s41467-021-21391-1.
[10] Alghamdi H, Maduabuchi C, Okoli K, Albaker A, Alanazi M, Alghassab M, et al.
From sunlight to power: enhancing 4E performance with two-stage segmented
thermoelectric generators in concentrated solar applications. J Clean Prod 2023;
429:139314. https://doi.org/10.1016/j.jclepro.2023.139314.
[11] C´
ozar IR, Pujol T, Lehocky M. Numerical analysis of the effects of electrical and
thermal congurations of thermoelectric modules in large-scale thermoelectric
generators. Appl Energy 2018;229:26480. https://doi.org/10.1016/j.
apenergy.2018.07.116.
[12] Yang W, Jin C, Zhu W, Li Y, Zhang R, Huang L, et al. Taguchi optimization and
thermoelectrical analysis of a pin n annular thermoelectric generator for
automotive waste heat recovery. Renew Energy 2024;220:119628. https://doi.
org/10.1016/j.renene.2023.119628.
[13] Yang W, Zhu W, Yang Y, Huang L, Shi Y, Xie C. Thermoelectric performance
evaluation and optimization in a concentric annular thermoelectric generator
under different cooling methods. Energies 2022;15:2231. https://doi.org/
10.3390/en15062231.
[14] Li G, Ying J, Zheng Y, Guo W, Tang Y, Ye C. Analytical design model for waste heat
thermoelectric generator and experimental verication. Energ Conver Manage
2022;252:115034. https://doi.org/10.1016/j.enconman.2021.115034.
[15] Zhao Y, Li W, Zhao X, Wang Y, Luo D, Li Y, et al. Energy and exergy analysis of a
thermoelectric generator system for automotive exhaust waste heat recovery. Appl
Therm Eng 2024;239:122180. https://doi.org/10.1016/j.
applthermaleng.2023.122180.
[16] Luo D, Yan Y, Chen W-H, Yang X, Chen H, Cao B, et al. A comprehensive hybrid
transient CFD-thermal resistance model for automobile thermoelectric generators.
Int J Heat Mass Transf 2023;211:124203. https://doi.org/10.1016/j.
ijheatmasstransfer.2023.124203.
[17] He M, Wang E, Zhang Y, Zhang W, Zhang F, Zhao C. Performance analysis of a
multilayer thermoelectric generator for exhaust heat recovery of a heavy-duty
diesel engine. Appl Energy 2020;274:115298. https://doi.org/10.1016/j.
apenergy.2020.115298.
[18] Pujol T, Jollyn IT, Massaguer E, Massaguer A, C´
ozar IR, Paepe MD. Design
optimization of plate-n heat sink with forced convection for single-module
thermoelectric generator. Appl Therm Eng 2023;221:119866. https://doi.org/
10.1016/j.applthermaleng.2022.119866.
[19] Chen W-H, Wang C-M, Saw LH, Hoang AT, Bandala AA. Performance evaluation
and improvement of thermoelectric generators (TEG): n installation and
compromise optimization. Energ Conver Manage 2021;250:114858. https://doi.
org/10.1016/j.enconman.2021.114858.
[20] Zhao Y, Lu M, Li Y, Wang Y, Ge M. Numerical investigation of an exhaust
thermoelectric generator with a perforated plate. Energy 2023;263:125776.
https://doi.org/10.1016/j.energy.2022.125776.
[21] Buononmo B, Cascetta F, Pasqua AD, Manca O. Performance parameters
enhancement of a thermoelectric generator by metal foam in exhaust automotive
lines. Therm Sci Eng Prog 2023;38:101684. https://doi.org/10.1016/j.
tsep.2023.101684.
[22] Li Y, Wang S, Zhao Y, Yue L. Effect of thermoelectric modules with different
characteristics on the performance of thermoelectric generators inserted in the
central ow region with porous foam copper. Appl Energy 2022;327:120041.
https://doi.org/10.1016/j.apenergy.2022.120041.
[23] He W, Guo R, Takasu H, Kato Y, Wang S. Performance optimization of common
plate-type thermoelectric generator in vehicle exhaust power generation systems.
Energy 2019;175:115363. https://doi.org/10.1016/j.energy.2019.03.174.
[24] Luo D, Wang R, Yu W, Zhou W. A numerical study on the performance of a
converging thermoelectric generator system used for waste heat recovery. Appl
Energy 2020;270:115181. https://doi.org/10.1016/j.apenergy.2020.115181.
[25] Yang W, Zhu W, Li Y, Zhang L, Zhao B, Xie C, et al. Annular thermoelectric
generator performance optimization analysis based on concentric annular heat
exchanger. Energy 2022;239:122127. https://doi.org/10.1016/j.
energy.2021.122127.
[26] Luo D, Wu Z, Yan Y, Cao J, Yang X, Zhao Y, et al. Performance investigation and
design optimization of a battery thermal management system with thermoelectric
coolers and phase change materials. J Clean Prod 2024;434:139834. https://doi.
org/10.1016/j.jclepro.2023.139834.
[27] Eiamsa-ard S, Wongcharee K, Promvonge P. Inuence of nonuniform twisted tape
on heat transfer enhancement characteristics. Chem Eng Commun 2012;199(10):
127997. https://doi.org/10.1080/00986445.2012.668724.
[28] Bucak H, Yılmaz F. The current state on the thermal performance of twisted tapes:
a geometrical categorisation approach. Chem Eng Process 2020;153:107929.
https://doi.org/10.1016/j.cep.2020.107929.
[29] Zhu W, Xu A, Yang W, Xiong B, Xie C, Li Y, et al. Optimal design of annular
thermoelectric generator with twisted tape for performance enhancement. Energ
Conver Manage 2022;270:116258. https://doi.org/10.1016/j.
enconman.2022.116258.
[30] Varun MO, Garg H Nautiyal, Khurana S, Shukla MK. Heat transfer augmentation
using twisted tape inserts: a review. Renew Sustain Energy Rev 2016;63:193225.
https://doi.org/10.1016/j.rser.2016.04.051.
[31] Karana DR, Sahoo RR. Performance assessment of the automotive heat exchanger
with twisted tape for thermoelectric based waste heat recovery. J Clean Prod 2021;
283:124631. https://doi.org/10.1016/j.jclepro.2020.124631.
[32] Demeke W, Ryu B, Ryu S. Machine learning-based optimization of segmented
thermoelectric power generators using temperature-dependent performance
properties. Appl Energy 2024;355:122216. https://doi.org/10.1016/j.
apenergy.2023.122216.
[33] Erro I, Aranguren P, Alzuguren I, Chavarren D, Astrain D. Experimental analysis of
one and two-stage thermoelectric heat pumps to enhance the performance of a
thermal energy storage. Energy 2023;285:129447. https://doi.org/10.1016/j.
energy.2023.129447.
[34] Alghamdi H, Maduabuchi C, Okoli K, Albaker A, Makki E, Alghassab M, et al.
Pioneering sustainable power: harnessing material innovations in double stage
segmented thermoelectric generators for optimal 4E performance. Appl Energy
2023;352:121885. https://doi.org/10.1016/j.apenergy.2023.121885.
[35] Ge M, Xuan Z, Liu X, Luo D, Wang Y, Li Y. Structural optimization of solar
thermoelectric generators considering thermal stress conditions. J Clean Prod
2023;428:139367. https://doi.org/10.1016/j.jclepro.2023.139367.
[36] Kim TY, Negash AA, Cho G. Waste heat recovery of a diesel engine using a
thermoelectric generator equipped with customized thermoelectric modules. Energ
Conver Manage 2016;124:2806. https://doi.org/10.1016/j.
enconman.2016.07.013.
[37] Lu X, Yu X, Qu Z, Wang Q, Ma T. Experimental investigation on thermoelectric
generator with non-uniform hot-side heat exchanger for waste heat recovery. Energ
Conver Manage 2017;150:40314. https://doi.org/10.1016/j.
enconman.2017.08.030.
[38] Ge M, Li Z, Zhao Y, Xuan Z, Li Y, Zhao Y. Experimental study of thermoelectric
generator with different numbers of modules for waste heat recovery. Appl Energy
2022;322:119523. https://doi.org/10.1016/j.apenergy.2022.119523.
[39] Negash AA, Choi Y, Kim TY. Experimental investigation of optimal location of ow
straightener from the aspects of power output and pressure drop characteristics of a
thermoelectric generator. Energy 2021;219:119565. https://doi.org/10.1016/j.
energy.2020.119565.
[40] Kim TY, Ewak J, Kim B-W. Energy harvesting performance of hexagonal shaped
thermoelectric generator for passenger vehicle applications: an experimental
approach. Energ Conver Manage 2018;160:1421. https://doi.org/10.1016/j.
enconman.2018.01.032.
[41] Wang T, Luan W, Liu T, Tu S-T, Yan J. Performance enhancement of thermoelectric
waste heat recovery system by using metal foam inserts. Energ Conver Manage
2016;124:139. https://doi.org/10.1016/j.enconman.2016.07.006.
[42] Zhang H, Nunayon SS, Jin X, Lai ACK. Pressure drop and nanoparticle deposition
characteristics for multiple twisted tape inserts with partitions in turbulent duct
ows. Int J Heat Mass Tran 2022;193:121474. https://doi.org/10.1016/j.
ijheatmasstransfer.2021.121474.
[43] Yang W, Zhu W, Du B, Wang H, Xu L, Xie C, et al. Power generation of annular
thermoelectric generator with silicone polymer thermal conductive oil applied in
automotive waste heat recovery. Energy 2023;282:128400. https://doi.org/
10.1016/j.energy.2023.128400.
[44] Yang W, Zhu W, Li Y, Xie C, Xiong B, Shi Y, et al. Global structural optimization of
annular thermoelectric generators based on a dual-nite-element multiphysical
model. Appl Therm Eng 2023;220:119797. https://doi.org/10.1016/j.
applthermaleng.2022.119797.
[45] Xuan Z, Ge M, Zhao C, Li Y, Wang S, Zhao Y. Effect of nonuniform solar radiation
on the performance of solar thermoelectric generators. Energy 2024;290:130249.
https://doi.org/10.1016/j.energy.2024.130249.
[46] Zhao X, Jiang J, Zuo H, Mao Z. Performance analysis of diesel particulate lter
thermoelectric conversion mobile energy storage system under engine conditions
of low-speed and light-load. Energy 2023;282:128411. https://doi.org/10.1016/j.
energy.2023.128411.
[47] Yang W, Xu A, Zhu W, Li Y, Shi Y, Huang L, et al. Performance improvement and
thermomechanical analysis of a novel asymmetrical annular thermoelectric
generator. Appl Therm Eng 2024;237:121804. https://doi.org/10.1016/j.
applthermaleng.2023.121804.
[48] Zhu W, Weng Z, Li Y, Zhang L, Zhao B, Xie C, et al. Theoretical analysis of shape
factor on performance of annular thermoelectric generators under different
thermal boundary conditions. Energy 2022;239:122285. https://doi.org/10.1016/
j.energy.2021.122285.
[49] Yin T, Li Z-M, Peng P, Liu W, Shao Y-Y, He Z-Z. Performance analysis of a novel
two-stage automobile thermoelectric generator with the temperature-dependent
materials. Appl Therm Eng 2021;195:117249. https://doi.org/10.1016/j.
applthermaleng.2021.117249.
W. Yang et al.
Applied Energy 363 (2024) 123047
14
[50] Vashistha C, Patil AK, Kumar M. Experimental investigation of heat transfer and
pressure drop in a circular tube with multiple inserts. Appl Therm Eng 2016;96:
11729. https://doi.org/10.1016/j.applthermaleng.2015.11.077.
[51] Luo D, Yan Y, Li Y, Yang X, Chen H. Exhaust channel optimization of the
automobile thermoelectric generator to produce the highest net power. Energy
2023;281:128319. https://doi.org/10.1016/j.energy.2023.128319.
[52] Yang Y, Wang S, Zhu Y. Evaluation method for assessing heat transfer
enhancement effect on performance improvement of thermoelectric generator
systems. Appl Energy 2020;263:114688. https://doi.org/10.1016/j.
apenergy.2020.114688.
W. Yang et al.
Article
Full-text available
Segmented thermoelectric generators (STEGs) provide an excellent platform for thermal energy harvesting devices because they improve power generation performance across a broad range of operating temperatures. Despite the benefit of direct thermal energy-to-electricity conversion, conventional STEG optimization approaches are unable to provide a systematic method for selecting the optimal multiple stacks of p- and-n-type thermoelectric materials (TEs) legs from a set of numerous TE materials. In this study, we propose a systematic optimization method based on machine learning to find the optimal STEG for performance maximization. A deep neural network (DNN) is trained using an initial dataset generated via the Finite Element Method (FEM), with inputs including the temperature-dependent properties of p- and n-type TE materials, the lengths of each segment, and external loads, as well as the corresponding performance outputs. The trained DNN captures the inherent nonlinear relationship between these inputs and outputs. The combination of a trained DNN and a genetic algorithm (GA) efficiently navigates a vast design space using 88 p-type and 70 n-type TE materials along with device factors. It formulates four stacked segment pairs in p- and n-leg TEGs, targeting new superior designs with enhanced power, efficiency, or both. The DNN is iteratively refined via active learning (AL) by incorporating new superior STEG designs to enhance prediction accuracy. The optimized STEGs exhibit power and efficiency that are 1.91 and 1.5 times higher, respectively, than the top designs from the initial training dataset composed of 157,916 STEGs. Furthermore, compared to conventional TEG designs without segmentation, our method discovered numerous high-performing STEG designs.
Article
Enhancing thermoelectric performance while minimizing exhaust back pressure is a crucial step in advancing the commercial viability of automotive thermoelectric generators. To achieve high overall performance in a thermoelectric generator, an annular thermoelectric generator equipped with circular pin fins is proposed. A comprehensive three-dimensional numerical model is established to accurately predict thermoelectric performance and thermomechanical behavior. Detailed multi-physics field distribution characteristics are analyzed. Using an L25 orthogonal array, we examine five influencing factors and their five levels: exhaust temperature, exhaust mass flow rate, fin height, fin diameter, and the number of fins. The Taguchi analysis suggests that exhaust temperature is the most influential factor in determining thermoelectric performance, followed by mass flow rate, fin height, fin diameter, and fin number. The optimal values for these parameters are 673 K, 30 g/s, 20 mm, 3 mm, and 420, respectively. Under the optimal design parameters, the net power reaches 34.11 W, representing an 18.7% increase compared to the original design. Moreover, a comparative study is conducted between plate fins and pin fins, showing that the pin fin-based thermoelectric generator exhibits a 5.83% increase in output power and a 4.82% increase in maximum thermal stress compared to the plate fin-based thermoelectric generator.
Article
This research investigates the Concentrated Solar Two-Stage Segmented Thermoelectric Generator (TSTEG) for efficient solar energy conversion. Through numerical optimization, the TSTEG's geometry parameters are fine-tuned to maximize performance under varying solar irradiances and heat transfer coefficients. A comprehensive 4E analysis evaluates key performance indicators: power output, exergy efficiency, entropy generation, CO2 savings per year, and dollar per watt. The results reveal optimized parameters, unlocking the TSTEG's full potential. The system achieves remarkable power generation of 41.12 W at a concentrated solar flux of 95 kWm−2, with a heat transfer coefficient of 4 kWm−2 K−1. Significant CO2 savings of 19.53 kgyr−1 are achieved. Furthermore, the peak exergy efficiency of 9.52% was attained at a concentrated solar flux of 20 kWm−2 and optimal semiconductor cross-sectional area of 0.01 mm2. Moreover, for optimal economic operation, the lowest dollar per watt value of 0.16 $/W is obtained at an optimal skutterudite amount of 92.59% under 40 kWm−2. These findings highlight the TSTEG's capacity for efficient power production and substantial carbon dioxide reduction, demonstrating its potential for sustainable and environmentally-friendly energy generation.
Article
Enhancing thermoelectric performance hinges on optimizing the geometry of thermoelectric legs. In this study, we present a novel asymmetrical annular thermoelectric generator (ATEG) in which the proportions of P-type and N-type legs are meticulously balanced. We construct a one-dimensional analytical model tailored to this ATEG. Utilizing this model, we derive the relationship governing thermal-electrical impedance matching in an asymmetrical ATEG and formulate a general expression for optimizing the asymmetry coefficient. We explore the influence of various thermal boundary conditions on optimal impedance matching, ideal annular leg parameters, and the optimal asymmetry coefficient. Our findings reveal that thermal boundary conditions significantly affect the optimal load ratio. Furthermore, in comparison to traditional ATEGs, our proposed asymmetrical ATEG with the optimized structure exhibits a remarkable 16.2% increase in output power while maintaining the same material volume. Additionally, we perform a three-dimensional numerical analysis of the asymmetrical ATEG using Comsol. Our research findings indicate that introducing the asymmetric structure leads to higher maximum thermal stress on the legs. Interestingly, the study of asymmetric thermal boundary conditions highlights that improving heat transfer between the ATEG and the cooler yields higher mechanical reliability compared to enhancing heat transfer between the ATEG and the heat source.