Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?

casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18


Chika Maduabuchi1
Department of Mechanical Engineering,
The Combined Impacts of Leg
University of Nigeria,
Nsukka, 410001 Enugu, Nigeria;
Department of Mechanical Engineering,
Geometry Configuration and
Federal University of Agriculture,
P.M.B. 2373, Makurdi, Nigeria Multi-Staging on the Exergetic
e-mail: Chika.maduabuchi.191341@unn.edu.ng

Sarveshwar Singh
Performance of Thermoelectric
Department of Electronics & Communication
Engineering, Modules in a Solar
Shobhit University,
250110 Meerut, Uttar Pradesh, India
e-mail: sarweshwarsingh@gmail.com
Thermoelectric Generator
Chigbogu Ozoegwu The performance of thermoelectric generators (TEGs) can be improved either by the adop-
Department of Mechanical Engineering, tion of multi-stage or tapered leg configuration. So far, a hybrid device that simultaneously
University of Nigeria, uses both multi-staging and tapered leg geometry to improve its performance has not been
Nsukka, 410001 Enugu, Nigeria conceived. Thus, we present a thermodynamic modeling and optimization of a two-stage
e-mail: chigbogu.ozoegwu@unn.edu.ng thermoelectric generator (TTEG) with tapered leg geometries using ANSYS 2020 R2 software.
The optimized parameters include the leg height, area, concentrated solar radiation, and
Howard Njoku external load resistance. First, the X-leg TEG only improves the performance of the trape-
Applied Renewable and Sustainable Energy
zoidal leg TEG below a leg height of 3 mm. Beyond 3 mm, the performance of both TEGs
Research Group,
become very similar. Long thermoelectric legs provide higher efficiencies, while short legs
Department of Mechanical Engineering,
generate maximum power densities. To obtain maximum efficiencies, the initial leg height of
University of Nigeria,
the thermoelectric legs, 1.62 mm, is increased by 517.28%, while the initial leg area,
Nsukka, 410001 Enugu, Nigeria;
1.96 mm2, is decreased by 64.29%. Also, the proposed TTEG with tapered legs (trapezoidal
Department of Mechanical Engineering Science,
and X-legs) improves the exergetic efficiency of the base case, single-stage rectangular leg
University of Johannesburg,
TEG, by 16.7%. Furthermore, the use of tapered leg TEGs, in single and multi-stage
Auckland Park, 2006 Johannesburg, South Africa
arrangements, reduces the exergy conversion index of conventional rectangular leg
e-mail: howard.njoku@unn.edu.ng
TEGs by 1.89% and 0.98%, respectively. Finally, the use of tapered legs and multi-stage
configurations increases the thermodynamic irreversibilities of conventional rectangular
leg TEGs, thus reducing their thermodynamic stability. [DOI: 10.1115/1.4051648]
Mkpamdi Eke1
Department of Mechanical Engineering, Keywords: solar energy conversion, thermoelectric generator, multi-stage design, tapered
University of Nigeria, leg geometry, geometry optimization, thermodynamic analysis, alternative energy sources,
Nsukka, 410001 Enugu, Nigeria energy conversion/systems, energy systems analysis, power (co-)generation, renewable
e-mail: mkpamdi.eke@unn.edu.ng energy

1 Introduction systems [7]. However, these devices are still characterized by low
efficiencies of 5% [8].
The dominant contemporary energy source over the past decades
An effective method of improving TEG performance is by
has remained as fossil fuel sources such as coal, oil, and natural gas
varying the geometry of the thermoelements. Also, the advent of
[1]. However, this has come at the price of environmental degrada-
additive manufacturing (three-dimensional printing) techniques
tion [2] and health hazards [3]. Also, the heavy reliance on fossil
has made the fabrication of leg geometries with complex shapes
fuel sources coupled with the exponential increase in world popula-
easy and possible [9], thus opening up a new field in thermoelectri-
tion suggests that fossil fuel sources will soon be exhausted [4,5].
city. Ali et al. [10] studied the effect of bi-tapered pin geometry,
Thus, exploring renewable energy sources such as solar energy
defined by a dimensionless shape parameter, on the first and
becomes inevitable.
second law efficiencies as well as the power output of a TEG.
Aside photovoltaic technology, a very lucrative method of con-
Fixed temperature ratios, θ = Tc/Th, of 0.3, 0.4, and 0.5 were used
verting solar energy directly into electricity is using thermoelectric
for the simulation. They discovered that the dimensionless shape
generators (TEGs). This is due to the remunerative advantages pro-
parameter significantly affected the second law efficiency while
vided by these devices such as noiseless operation, relatively low
slightly affecting the first law efficiency. Liu et al. [11] concluded
cost, zero environmental emissions, solid-state operation, suitability
that under isoflux conditions, thermoelectric (TE) legs with
for integrated systems, long lifespan, no moving parts, and low
varying leg areas (trapezoidal, parabolic, and exponential) always
maintenance requirements [6]. They have found various applica-
outperformed those with constant cross sections (rectangular/
tions such as power generation in severe operating conditions,
cuboid legs). Shittu et al. [12] argued that geometric optimization
waste heat recovery systems, integrated power systems, small-scale
could reduce the adverse impacts of varying meteorological condi-
power generation in electronic equipment, and solar powered
tions on TEG performance. Thermodynamic analysis of a TEG was
also carried out by Lamba and Kaushik [13] in which they investi-
gated the impacts of leg geometry configuration and Thomson
1
Corresponding authors. effect on the power output and efficiency of the device. Their
Contributed by the Advanced Energy Systems Division of ASME for publication in results showed that, at θ = 0.5, a trapezoidal shaped TEG gave a
the JOURNAL OF ENERGY RESOURCES TECHNOLOGY. Manuscript received January 27,
2021; final manuscript received June 23, 2021; published online July 16, 2021. 2.32% and 2.31% increase in energy and exergy efficiency of con-
Assoc. Editor: Guangdong Zhu. ventional TEGs, respectively. Ferreira-Teixeira and Pereira [14]

Journal of Energy Resources Technology Copyright © 2021 by ASME APRIL 2022, Vol. 144 / 041303-1
om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18
Fig. 1 Proposed TEG systems comprising two and four thermoelements for the single-stage
and two-stage TEGs, respectively. (a) System 1, (b) system 2, (c) system 3, (d) system 4,
(e) system 5, and (f ) system 6. Systems 1, 2, and 3 are the single-stage TEGs with rectangular,
trapezoidal, and X-legs, respectively. Systems 4, 5, and 6 are the two-stage TEGs with rectan-
gular, trapezoidal, and X-legs, respectively. The leg height and area of the rectangular, trape-
zoidal, and X-legs are the same.

showed that both cubic and cylindrical leg configurations offered during the study. Wang et al. [16] showed that for a fixed hot junc-
identical TEG performance and that utilizing more thermocouples tion temperature of 500 K and a draft angle of 10 deg, the X-leg
improved device overall performance. Ibeagwu [15] numerically TEG improved the power output of the conventional rectangular
modeled the performance of five leg geometries on COMSOL MULTI- leg TEG by 4.57% with the thermal stresses decreasing with
PHYSICS. The geometries modeled were the rectangular, trapezoidal, increasing draft angles. Niu et al. [17] showed that variable leg
I, Y, and X cross-sectional areas. It was discovered that varying area TEGs performed better than conventional rectangular leg
cross-sectional areas significantly influence the performance of con- area TEGs. The shapes considered were the conventional cuboid
ventional rect-leg with X-leg being the most efficient and showing a and hexahedron shapes. Despite the innovative findings made by
19.13% increase in power density of the rect-leg. Again, fixed these authors, an underlying assumption employed in the numerical
values of Th and Tc of 420 K and 300 K, respectively, were utilized model of these papers is that the same hot and cold junction

041303-2 / Vol. 144, APRIL 2022 Transactions of the ASME


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18
temperatures were applied to the rectangular leg and tapered leg TEGs are also studied. The results provided by this optimization
TEGs, respectively. This violates the experimental findings of study should be able to provide TEG manufactures and designers
Fabián-mijangos et al. [18], who reported that under the same with a more realistic and comprehensive perspective pertaining
heat flux, the trapezoidal leg TEG generated a higher temperature tapered leg/multi-stage TEG performance.
difference relative to the rectangular leg TEG. Thus, the current
information provided by these papers needs urgent improvements.
This will be done using isoflux boundary conditions rather than iso- 2 System Description and Materials
thermal ones as used by these cited papers.
Despite the improvements obtained from varying the geometry of Figure 1 shows the proposed systems considered in this study.
TE pins, several researchers went further to investigate the effect of Figures 1(a)–1(c) show systems 1–3 which comprise the single-
utilizing multi-stage TEGs. This is achieved by placing a TEG in stage setups while Figs. 1(d)–1(f ) illustrate systems 4–6 which
series or parallel with another TEG in order to improve the overall are the equivalent double-stage setups. The leg geometries consid-
system performance [19]. Cheng et al. [20] showed that the utiliza- ered are the traditional rectangular leg, trapezoidal leg, and the
tion of cascaded TEGs greatly improved the performance of TE recently introduced X-leg as seen in Figs. 1(a)–1(c), respectively.
devices at large temperature gradients. Zhang et al. [21] found that The standard leg height and area of the rectangular leg are obtained
a two-stage TEG was feasible in recovering waste heat from a from that of a contemporary module, as reported in Ref. [30]. The
solid oxide fuel cell system. Cheng et al. [22] reported that when trapezoidal leg geometry is modeled by maintaining the rectangular
incorporated in a hypersonic vehicle, a single-stage TEG provided leg area at one end, while using half of it at the other end. The X-leg
a higher power output density of 16.53 kW/m2, while a two-stage on the other hand is modeled with the rectangular leg area at both
TEG yielded a higher conversion efficiency of 10.78% for the ends, while using half of it at the mid-section. The initial leg
same inflow temperature. Lee et al. [23] studied the combined height and area are 1.6 mm and 1.4 × 1.4 mm2, respectively. The
effects of leg geometry, material segmentation, and two-stage leg dimensions of the single-stage setups are applicable to that of
arrangement on the electrical and mechanical performance of a the two-stage setups. The solar flux from the sun is magnified
TEG using ANSYS 19.1 software and found that the best performance using a compound parabolic concentrator. A surface selective
was obtained from a single-stage arrangement with segmented cylin- absorber (SSA) is placed at the TEG hot junction to convert the inci-
drical legs. Liu et al. [24] used a multi-objective optimization method dent concentrated light energy from the sun into sensible thermal
to optimize the power output and conversion efficiency of a two- energy. To necessitate a temperature gradient for power generation,
stage TEG and found that a weight factor of 0.5 was sufficient to the TEG cold junction is cooled by water. A convective film coef-
simultaneously improve device power output and conversion effi- ficient is used to specify the TEG water cooling.
ciency. Cheng et al. [25] concluded that multi-stage TEGs are For the material assignments in a conventional TEG, the hot and
more favorable for high-temperature applications compared to cold junction ceramic plates are specified as alumina. This material
single-stage setups. Zhang et al. [26] used a Latin hypercube sam- maintains the temperature gradient across both junctions, owing to
pling method to evaluate the performance of a two-stage TEG and its relatively low thermal conductivity. The conductor pads are spe-
concluded that the TEG cold side convective heat transfer coefficient cified as copper, due to its high electrical conductivity, thus allow-
and hot junction area contributed the least in determining device ing the continuous flow of current throughout the TEG. A solder
performance. Kanimba et al. [27] showed that under a fixed hot material, Sn–Pb (60–40%), is used to join the thermoelectric legs
and cold junction temperature of 973 K and 73 K, respectively, a to the copper pads. It is also used to reduce the thermal stress
power output of 505 W was obtained from a three-stage TEG. Sun levels in the legs [31]. The thermal and electrical contact between
et al. [28] utilized Shannon’s entropy method in selecting a suitable the solder material and the TE legs is assumed negligible [32].
Pareto front solution in the optimization of a two-stage TEG and con- Finally, the thermoelectric legs are made of pure bismuth-telluride
cluded that the Nadir model generated the least specific power material. This choice was made due to its commercial availability
output. Guo et al. [29] obtained 11.8%, 17.5%, and 17.7% increase and contemporary usage. The temperature-dependent material prop-
in the electric efficiency, exergy efficiency, and power output, erties of bismuth-telluride are shown in Fig. 2. The other TEG mate-
respectively, of a high-temperature proton exchange membrane rial properties are shown in Table 1.
fuel cell when a two-stage TEG was incorporated. These papers
establish that two-stage TEGs perform better than single-stage
TEGs. However, scarcely has anything been done on a two-stage 3 Methodology
TEG that employs tapered leg geometry configuration. The only
3.1 Analysis. The steady-state thermal-electric equations gov-
paper which attempted this was Ref. [23]. However, their discourse
erning thermoelectricity are [12,13]
was limited to a first law/energy analysis of the various TEG models.
The exergetic implication of utilizing variable area leg geometries    
and multi-stage configuration was not discussed in their report. ∇ · (k ∇ T) + J 2 ρ − τ J · ∇ T = 0 (1a)
Also, nothing was said pertaining the geometry optimization of the    
thermoelements. Moreover, the temperature dependence of the ther-  1  1 
∇· ∇ϕ + ∇ · α∇T = 0 (1b)
moelectric materials was not considered. Furthermore, the recently ρ ρ
introduced X-leg geometry [15,16] was not considered in their anal-
ysis. Finally, the unrealistic isothermal boundary condition was where α, ρ, k, and τ are the temperature-dependent Seebeck coeffi-
applied in their analysis of the single/multi-stage and rectangular/ cient, electrical resistivity, thermal conductivity, and Thomson

tapered leg TEGs. coefficient, respectively. J is the current density vector and ϕ is
To this end, this research paper seeks to fill up these research gaps the scalar potential of the electric field.
by conducting a second law thermodynamic optimization of a The heat input and output at the hot and cold junctions of the
multi-stage TEG with tapered leg geometry configuration under TEG are given as [8,37]
actual isoflux boundary conditions. Furthermore, a geometry opti-
mization is used to determine the optimum leg height and area for Qh = αTh I + KΔT − 0.5I 2 R (2a)
maximum power or efficiency in the tapered leg TEGs. The
temperature-dependent material properties of the thermoelectric Qc = αTc I + KΔT + 0.5I 2 R (2b)
materials are also considered. The optimum solar flux and external
load resistance for optimum performance are also discussed. The where Th and Tc are the hot and cold junction temperatures, respec-
effects of tapered leg geometry and multi-stage arrangements on tively. I is the current generated, K is the thermoelectric leg thermal
the exergetic efficiency and thermodynamic irreversibilities of the conductance, R is the electrical resistance to current flow offered by

Journal of Energy Resources Technology APRIL 2022, Vol. 144 / 041303-3


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18
Fig. 2 Temperature-dependent material properties: (a) Seebeck coefficient, (b) Thomson coefficient, (c) thermal conductivity,
and (d) electric resistivity [33]

the legs, and ΔT is the temperature gradient across the TEG hot and The heat inflow at the solar thermoelectric generator (STEG) hot
cold junctions. junction is given as [41]
Analyzing the TEG as a thermodynamic heat engine, the power
output and thermal efficiency are defined as [38] Qf = ηopt ηa τg ψAs (5)
where ηopt is the optical efficiency of the solar concentrator, ηa is the
Pte = Qh − Qc (3) efficiency of the SSA placed at the TEG hot junction, τg is the trans-
missivity of the concentrator glass cover, ψ is the concentrated solar
radiation, in suns, emanating from the solar concentrator, and As is
Qc the surface area of the SSA.
ηte = 1 − (4) Similarly, the exergy inflow at the STEG hot junction is obtained
Qh using Petela’s theory of exergy radiation which is defined as [42]
     
The TEG generates maximum power when the intrinsic resis- 4 Ta 1 Ta 4
tance offered by the TE legs matches up with that of the external Exf = Qf 1 − + (6)
3 Ts 3 Ts
load the TEG is supplying current to [39,40].

Table 1 Other TEG material properties

Specific heat capacity Density Length Thermal conductivity Electric Seebeck


Material (J kg−1 K−1) (kg m−3) (mm) (W m−1 K−1) resistivity (Ω m) coefficient (V K−1) References

Alumina 900 3900 0.8 25 – – [15]


Solder paste 210 7240 0.1 37.8 5 × 10−8 – [34]
Copper 800 3970 0.2 401 1.72 × 10−8 – [35,36]
Bismuth-telluride 154.4 7740 1.62 f(T ) f (T) f (T) [30,33]

041303-4 / Vol. 144, APRIL 2022 Transactions of the ASME


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18
Table 2 Simulation parameters

Parameter Symbol Unit Value References

Glass transmissivity τg – 0.95 [49]


Absorber efficiency ηa – 0.95 [41]
Selective surface effective emittance ɛs – 0.05 [41]
Stefan–Boltzmann constant σSB W m−2 K−4 5.67 × 10−8 [41]
Wind speed v m s−1 1 [39]
Ambient temperature Ta K 295.15 [39]
Optical efficiency ηopt – 0.95 [39]
Sun effective temperature Ts K 5777 [50]

where Ts is the sun effective blackbody temperature and Ta is the (3) An external load resistance is connected across the pin
temperature of the ambient environment. terminals.
Thus, the system energy and exergy efficiencies are given as, (4) Radiative and convective losses are specified at the TEG hot
respectively [43,44], junction.
(5) All external TEG surfaces are assumed perfectly insulated.
Pte
ηen = (7)
Qf 3.3 Validation. It is a customary practice to validate the
numerical model by using it to reproduce the works of previous
Pte authors on the same subject [51,52]. This is done so as to vet the
ηex = (8)
Exf accuracy and reliability of the present model. If the present model
gives results that are similar to those obtained by previous
Maduabuchi et al. [45,46] recently introduced a way of estimat- authors, then it can be trusted in performing the present optimiza-
ing the irreversibilities in a STEG using thermodynamic analysis, tion study. To this effect, we utilize the present model in reprodu-
which is given as cing the works of Refs. [18,53]. The single-stage TEGs are
validated with the experimental (expt.) results of Ref. [18], while
Exd = Exf − Pte (9)
the double-stage TEGs are validated with reports of Ref. [53].
where Exd is the amount of exergy destroyed. The results of the validation study are depicted in Fig. 3. First,
Similarly, the entropy generation is defined as Fig. 3(a) shows that the temperature gradient across the rectangular
leg and trapezoidal leg TEGs increases with the applied heat flux. It
Sgen =
Ed
(10) is also demonstrated that for the same heat flux variation, the trape-
Ta zoidal leg TEG generates a higher temperature gradient than the rec-
tangular leg TEG. These experimental results disprove the results of
Furthermore, a new exergy performance parameter called the previous optimization studies that assumed the same temperature
exergy conversion index has been proposed which, unlike previous gradient across the trapezoidal and rectangular leg TEG during
ones [47,48], provides an easy overview of the system exergy per- their analysis. Also, Fig. 3(b) shows that the conversion efficiency
formance. This is because the range of values necessary for profit- of a two-stage TEG increases as the temperature gradient is
able exergy evaluation is when it is greater than unity. increased. Overall, both plots reveal that a fair agreement exists
Exf between the present and previous results, thus declaring the
μex = (11) present model accurate and reliable.
Ed
The radiative and convective losses at the STEG hot junction are
specified as follows [39]: 4 Results and Discussions
Tsky = 0.0552Ta1.5 (12) 4.1 Effect of Leg Geometry Parameters
4.1.1 Effect of Leg Height. The dimensions of the thermoelec-
h = 5.82 + 4.07v (13) tric legs, comprising the leg height and area, are crucial parameters
that affect the device overall performance. This is because they
where v is the wind speed. determine the rate of heat absorption and rejection at the TEG hot
Table 2 illustrates the parameters used during the simulation. and cold junctions, respectively. Thus, this sub-section is dedicated
to studying the effects of these parameters on the device overall per-
formance. The leg height is varied from 1 mm to 10 mm, while the
leg area ranges from 0.7 mm2 to 2.59 mm2. These ranges are
3.2 Computational Route and Boundary Conditions. For
selected in order to go below and above the reference leg heights
reasons earlier explained in Sec. 1, the commercially obtainable
and areas of 1.62 mm and 1.96 mm2, respectively, as specified by
finite element solver, ANSYS 2020 R2 software, is used in this
the TEG manufacturer [30], thus encompassing both short/long
study. First, the three-dimensional computer aided design models
and thin/thick leg applications. This will give the TEG designer
are developed in AUTODESK INVENTOR 2021 software and are
and manufacturer a broader view of the impact of leg height and
imported directly to ANSYS Workbench platform. Then, the
area, consequently facilitating their decision-making process.
thermal and electric solvers are coupled in the software interface
Figures 4 and 5 show the results of the leg height and area optimi-
and are utilized in solving the steady-state thermoelectric field equa-
zation, respectively.
tions. The following boundary conditions are applied to the TEG
The variation of the temperature gradient and power output gen-
model in order to represent a real-life scenario:
erated per unit leg volume is shown in Figs. 4(a) and 4(b), respec-
(1) Unless otherwise stated, a fixed concentrated solar radiation tively. The temperature gradient increases as the leg height is
of 10 suns is maintained at the TEG hot junction. increased, while the power output density decreases. This contrast-
(2) A convective heat transfer coefficient of magnitude, ing relationship is due to the fact that increasing the leg height
500 W/m2 K, is applied at the TEG cold junction. increases the heat absorption and rejection rate at the TEG hot

Journal of Energy Resources Technology APRIL 2022, Vol. 144 / 041303-5


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18
Fig. 3 Validation results: (a) single-stage rectangular and trapezoidal leg TEGs [18] and (b) two-stage rectangular leg
TEG [53]

and cold junctions, respectively, while increasing the leg electrical obtained from a system while providing means to measure the
resistance and volume. The former increases the temperature gradi- system irreversibilites which reduce the thermodynamic perfor-
ent, while the latter reduces the electrical power output per unit leg mance. These will be shown in subsequent plots. Second, the order
volume. Also, the increase provided in the temperature gradient of increase in the efficiencies of the systems are 1, 4/2/3, and 5/6,
reduces with increasing leg heights. For instance, in system 6, as respectively. It must be emphasized that the highest efficiencies
the leg height is increased from 1 mm to 2 mm, the temperature gra- obtained from systems 4/2/3 and 5/6 are heavily influenced by the
dient increases by 57.43%, while for a leg height increase from leg height. Thus, for short leg height applications, requiring
9 mm to 10 mm, the temperature gradient rises by just 2.94%. maximum efficiencies, system 3 will be preferred to systems 2 and
This shows that the TEG manufacturer must strike an economic 4. However, if long legs are used, the same efficiencies will be pro-
balance in utilizing very long legs in TEGs. Overall, for the same duced by the three systems. Similarly, for short legs, system 6 is
amount of concentrated solar radiation, as the leg height is recommended above system 5. But, when long legs are utilized, the
increased, varying temperature gradients are obtained from the same efficiencies are obtained from both systems. Lastly, a very
diverse systems studied. First, systems 5 and 6 generate similar tem- important observation is made, in that as the leg height is increased,
peratures at longer leg heights, with the latter outperforming the the power output density decreases, as shown in Fig. 4(b), while
former at shorter leg heights. Second, systems 2, 3, and 4 yield the current plots show that the energy/exergy efficiencies increase
similar temperatures at longer leg heights, while recording accordingly. This is due to the independence of the efficiencies on
varying values at shorter leg heights, with the highest values the leg volume and its high dependence on the temperature gradient.
yielded by systems 3, 2, and 4, respectively, in that order. Lastly, Thus, the TEG manufacturer must make an important decision on
the conventional system 1 gives the lowest temperature gradients. whether to design the TEG for maximum power (short legs) or
These findings violate the reports of previous optimization maximum efficiency (long legs). However, since this study utilizes
studies, which assumed the same hot and cold junction temperatures exergy analysis, we design the device for maximum efficiency.
in the comparative study of TEGs with diverse leg geometries. In Thus, an optimum leg height of 10 mm is chosen.
addition, they also improve the work of Ibeagwu [15], who con- Then, Figs. 4(e) and 4(f ) show the effects of leg height on the
cluded that the X-leg outperformed the trapezoidal leg under all system irreversibilites and entropy generation per unit volume.
conditions. However, their study was not broad enough to accom- The system irreversibilites and entropy generation are thermody-
modate a detailed geometry optimization. namic parameters that are associated with the second law of state-
Furthermore, Fig. 4(b) shows that the order of increase in the ment. They give perspective on the magnitude of all the
power output densities depends on the leg height used. For instance, combined factors contributing to exergy destruction within the
for leg heights below 3 mm, system 2 generates higher power system. In a solar TEG, such factors comprise the thermal and elec-
output densities than system 3 while system 5 outpaces system trical resistance offered by the thermoelectric legs, thermal losses at
6. Beyond 3 mm, the systems 2/3 and 5/6 generate similar power the STEG hot junction, imperfect electrical and thermal contact
output densities. Besides, at a critical leg height, the single-stage renitences, thermal limitations of solid-state materials, and inelastic
TEG generates equal power output density with its double-stage distortions at elevated temperatures. It is seen that as the leg height
counterpart. This is because of the increased electrical resistance is increased, the system irreversibilites and entropy generation den-
offered by the two-stage TEG coupled with its larger material sities begin to decrease exponentially. This is the reason for the rise
volume owing to a larger number of legs. The optimum leg in energy/exergy efficiencies reported in Figs. 4(c) and 4(d), respec-
height for systems 2/3 and 5/6 is 5 mm, while that of system 1/4 tively. Thus, shorter legs increase exergy destruction in a STEG
is 10 mm. This indicates that the two-stage variable leg area system, while longer legs are crucial to obtaining increased thermo-
TEGs reduce the critical leg height of the traditional rectangular dynamic stability.
leg TEG by 100%, thus saving more material and ultimately cost. Finally, we consider the impact of leg height on the STEG exergy
Meanwhile, Figs. 4(c) and 4(d ) show the trend obtained in the conversion index. This is a new exergy performance coefficient that
energy/exergy efficiencies as the leg height is increased. First, the was introduced by the authors for the sake of simplifying exergy
exergy efficiency is slightly higher than the energy efficiency. This analysis. We define it as the ratio between the exergy inflow, at
is in perfect accordance with the laws of thermodynamics [38]. the STEG hot junction, and the exergy destroyed due to the gener-
This is because, unlike an energy analysis, an exergy analysis mea- ation of system irreversibilites. Thus, the ratio for reasonable exergy
sures the highest possible theoretical performance that can be performance is that the exergy conversion index must be greater

041303-6 / Vol. 144, APRIL 2022 Transactions of the ASME


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18

Fig. 4 Effect of leg height on (a) temperature gradient, (b) power output density, (c) energy efficiency,
(d) exergy efficiency, (e) irreversibilities density, ( f ) entropy generation density, and (g) exergy conversion
index. Systems 1, 2, and 3 are the single-stage TEGs with rectangular, trapezoidal, and X-legs, respectively.
Systems 4, 5, and 6 represent the two-stage TEGs with rectangular, trapezoidal, and X-legs, respectively. The
thermoelectric leg length is varied from 1 mm to 10 mm while maintaining a leg area of 1.96 mm2. OTEGs
stand for the one-stage TEGs, while TTEGs are the two-stage TEGs.

Journal of Energy Resources Technology APRIL 2022, Vol. 144 / 041303-7


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18

Fig. 5 Effect of leg area on (a) temperature gradient, (b) power output density, (c) energy efficiency, (d) exergy
efficiency, (e) irreversibilities density, ( f ) entropy generation density, and (g) exergy conversion index. Systems
1, 2, and 3 are the single-stage TEGs with rectangular, trapezoidal, and X-legs, respectively. Systems 4, 5, and 6
represent the two-stage TEGs with rectangular, trapezoidal, and X-legs, respectively. The leg area is varied from
0.7 mm2 to 2.59 mm2 while maintaining an optimum leg height of 10 mm. OTEGs stand for the one-stage TEGs,
while TTEGs are the two-stage TEGs.

041303-8 / Vol. 144, APRIL 2022 Transactions of the ASME


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18
than one. This is because, if it is less than or equal to one, it means increase in the efficiencies of the diverse systems become 1, 2/3, 4,
that the system has a higher rate of exergy destruction compared to and 5/6, respectively.
the exergy coming in from the heat source. Thus, much exergy is Then, Figs. 5(e) and 5(f ) show the effect of leg area on the irre-
being wasted. To this end, Fig. 4(g) shows the variation of the versibilities and entropy generation densities of the diverse systems.
exergy conversion index in the one-stage thermoelectric generators As expected, the system irreversibilities reduce as the leg area is
(OTEGs) and two-stage thermoelectric generators (TTEGs). These increased. This is due to the higher rate of exergy destruction asso-
parameters were plotted on two different y-axes due to the large dif- ciated with using thick areas. This also corroborates the results of
ference between them, which ultimately affected the result presen- the energy/exergy efficiencies earlier discussed, since an increase
tation. First, the exergy conversion index of a TEG increases with in the thermodynamic efficiencies is a consequence of reduced
the stage height. In addition, while the exergy efficiency increases system irreversibilites. It is also seen that the highest irreversibilities
with leg height, the exergy conversion index on the other hand is and entropy generation densities are obtained from system 4/5, fol-
maximized at an optimum leg height. This tells that the current irre- lowed by system 4, next to systems 2/3, and finally, system 1. This
versibility analysis is not sufficient to maximize exergy conversion tells that the adoption of variable leg geometries/multi-staging
in thermodynamic devices. The leg height range specified was not comes at the expense of the traditional systems’ thermodynamic
enough to maximize the exergy conversion index of system 1, as stability.
it kept increasing even beyond 10 mm. However, a careful observa- Finally, Fig. 5(g) goes beyond just estimating the system irrevers-
tion of the trend shows that the increase in its exergy conversion ibilities but measures these irreversibilites with respect to the solar
index begins to decrease with increasing leg heights. This tells exergy input. We see that the exergy conversion index is optimized
that its optimum leg height exceeds 10 mm. Besides, the optimum at an optimum leg area different from that for maximum exergy effi-
leg lengths for maximum exergy conversion in systems 2/3/4, 5, ciency. The optimum leg area for maximum exergy conversion in
and 6 are 7 mm, 4 mm, and 3 mm, respectively. This tells that system 1 is 1.33 mm2. Since our focus was to optimize the
the optimum leg height for maximum exergy conversion in a single- exergy efficiency, the area range specified was even inadequate to
stage trapezoidal/X-leg TEG and double-stage rectangular leg TEG optimize the exergy conversion indices of systems 2–6, respec-
is the same. Typical numerical values show that system 6 reduces tively. This tells that while maximum exergy efficiency is obtained
the optimum leg height in systems 2/3/4 and 5 by 57.14% and at 0.7 mm2, the exergy conversion can only be maximized, that is
25%, respectively. Also, system 6 reduces the optimum leg height the full conversion of input solar exergy to useful exergy output,
of systems 2/3/4 by 42.86%. Therefore, the optimum leg height when thicker legs, exceeding 2.6 mm2, are used.
decreases as the number of stages in the TEGs is increased. This
also infers that the use of variable leg geometries in multi-stage
TEGs reduces the amount of material required to maximize 4.2 Effect of Concentrated Solar Radiation. The concen-
device exergy conversion. trated solar radiation from the solar concentrator is a major
thermal parameter that influences the overall performance of a
4.1.2 Effect of Leg Area. For an optimum leg height of 10 mm, solar TEG. This is because it determines the magnitude of the hot
the effects of thermoelectric leg area on the TEG temperature gradi- and cold junction temperatures in the device, thus determining the
ent, power output density, energy efficiency, and exergy efficiency temperature gradient and the overall device performance. In addi-
are captured in Figs. 5(a)–5(d), respectively. It is noticed that all tion, the thermal limitations of bismuth-telluride necessitate a
these parameters decrease as the leg area is increased. This is due study on this crucial thermal parameter. Thus, in this section, it is
to the fact that as the leg area is increased, the rate of heat transfer, varied from 5 to 50 suns (1000 W m−2 = 1 sun). This is done in
by conduction, from the leg hot junction to cold junction increases. order to encompass both low and high solar flux intensities. Our
Thus, meaning that less heat will be retained at both junctions nec- major aim is to maximize the system exergy efficiency.
essary to maintain a high temperature gradient. Hence, the decrease Figures 6(a) and 6(b) show the variation of the temperature gra-
in temperature gradient as the leg area is increased. This information dients and power output densities obtained from the diverse systems
tells the TEG designer and manufacturer that maximum system per- studied. We see that both parameters increase as the concentrated
formance can be obtained while economizing the material volume. solar radiation intensity is increased. This is due to the increased
Also, the temperature gradient obtained across systems 2/3/4 and heat absorption at the TEG hot junction. Furthermore, the increase
5/6 is very similar, with system 1 generating the least temperature in both parameters begins to decline even as the concentrated solar
gradients. radiation intensity is further increased. The solar flux range speci-
Unexpectedly, Fig. 5(b) shows that the highest power output den- fied was not even enough to maximize the power output density.
sities are obtained from systems 2/4, although systems 5/6 generate This is because our focus was not to maximize the power output
the highest temperature gradients as seen in Fig. 5(a). This can be density but the exergy efficiencies. Also, the order of increase in
explained by considering that increasing the leg area results in an the temperature gradients of the various systems is different from
increase in material volume. Furthermore, the legs in systems 2/3 that of the power output densities. For instance, in Fig. 6(a), the
are twice that of systems 5/6. This results in a substantial increase order of increase is systems 5/6, 2/3/4, and 1, respectively, while
in the power output density of the former relative to the latter, irre- in Fig. 6(b), the order of increase becomes systems 2/3, 1, and 4/
spective of the higher temperature gradients generated by the latter. 5/6, respectively. The higher power output densities obtained
For similar reasons, below a critical area of 1.75 mm2 and from systems 2/3 are because of the lesser material volume it occu-
1.12 mm2, system 1 generates higher power output densities than pies compared to all other systems. Also, system 1 offers a lesser
systems 4 and 5/6, respectively. Beyond these critical values, the material volume compared to systems 4/5/6. These facts render
latter systems generate higher power output densities compared to the higher temperature gradients developed by the systems 4/5/6
the former. compared to 1/2/3 inconsequential. However, a further observation
Due to the independence of the energy/exergy efficiencies on the of Fig. 6(b) reveals that at a critical solar flux of 50 suns, the power
leg volume, Figs. 5(c) and 5(d) show that systems 5 and 6 generate output densities obtained from systems 2/3/1 becomes the same. For
the highest values. This is also a function of the relatively higher solar fluxes below this critical value, systems 2/3 offer higher values
temperatures obtained from these systems. This is followed by compared to system 1. Finally, below and above a critical value of
system 4, which outperforms systems 2/3 only above a critical 25 suns, the power output densities from systems 4 become lower
value of 1.25 mm2. Beyond this critical area, the efficiencies and higher than that of systems 5/6 combined, respectively.
obtained from systems 2/3/4 are very similar. Finally, the least effi- On the other hand, Figs. 6(c) and 6(d) show the behavior of the
ciencies are obtained from the traditional system 1. However, since energy/exergy efficiencies with the concentrated solar flux, respec-
our study embraces a thermodynamic/exergy analysis, the leg area tively. It is seen that there exists an optimum solar flux at which
of 0.7 mm2 is selected as the optimum leg area. Thus, the order of maximum energy/exergy efficiencies are obtained from the various

Journal of Energy Resources Technology APRIL 2022, Vol. 144 / 041303-9


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18

Fig. 6 Effect of concentrated solar radiation intensity on (a) temperature gradient, (b) power output density,
(c) energy efficiency, (d) exergy efficiency, (e) irreversibilities density, ( f ) entropy generation density,
and (g) exergy conversion index. Systems 1, 2, and 3 are the single-stage TEGs with rectangular, trapezoidal,
and X-legs, respectively. Systems 4, 5, and 6 represent the two-stage TEGs with rectangular, trapezoidal, and
X-legs, respectively. The concentrated solar radiation intensity is varied from 5 to 50 suns for a fixed
optimum leg height and area of 10 mm and 0.7 mm2, respectively. OTEGs stand for the one-stage TEGs, while
TTEGs are the two-stage TEGs.

041303-10 / Vol. 144, APRIL 2022 Transactions of the ASME


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18

Fig. 7 Effect of external load resistance on (a) power output density, (b) energy efficiency, (c) exergy effi-
ciency, (d ) irreversibilities density, (e) entropy generation density, ( f ) irreversibilities and entropy generation
densities for system 1, and (g) exergy conversion index. Systems 1, 2, and 3 are the single-stage TEGs with
rectangular, trapezoidal, and X-legs, respectively. Systems 4, 5, and 6 represent the two-stage TEGs with rec-
tangular, trapezoidal, and X-legs, respectively. The concentrated solar radiation intensity is varied from 5 to
50 suns for a fixed optimum leg height and area of 10 mm and 0.7 mm2, respectively. OTEGs stand for the
one-stage TEGs, while TTEGs are the two-stage TEGs.

Journal of Energy Resources Technology APRIL 2022, Vol. 144 / 041303-11


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18
systems studied. The optimum solar flux required to maximize the rectangular leg TEG. In addition, multi-staging reduces the
efficiencies varies across the single and double-stage systems. For optimum load for maximum power in a single-stage, rectangular/
instance, the optimum concentrated solar radiation intensities variable leg area, TEG. Also, beyond a critical load resistance of
required to maximize the energy/exergy efficiencies in systems 0.7 Ω, the power output densities generated by systems 2/3 exceed
1/2/3 are 15 suns, while that required in systems 4/5/6 are 20 suns. that of system 1. However, above a critical load of 0.8 Ω, systems
This means that it will be more expensive to maximize the energy/ 5/6 outperform system 4. This means that there exists a critical
exergy efficiencies in a two-stage system compared to a single-stage load resistance beyond which the power output density from a two-
system, since a higher solar flux comes at a higher concentration ratio stage TEG becomes higher than that of a two-stage TEG.
and ultimately cost. Finally, the maximum energy/exergy efficien- For the energy/exergy efficiencies, Figs. 7(b) and 7(c), respec-
cies obtained from systems 1, 2/3, 4, and 5/6 are 6.13%/6.58%, tively, show that the optimum load resistances at which
6.24%/6.69%, 6.86%/7.36%, and 6.98%/7.5%, respectively. maximum energy/exergy efficiencies are obtained from systems
Also, Figs. 6(d) and 6(e) show that the system irreversibilities 1, 2/3, 4, and 5/6 are 0.8 Ω, 1.5 Ω, 0.7 Ω, and 1.3 Ω, respectively.
and entropy generation densities in the various systems considered These optimums correspond to maximum energy/exergy efficien-
increase with the concentrated solar radiation intensity emanating cies of 6.03%/6.47%, 5.93%/6.36%, 6.9%/7.41%, and 7.03%/
from the solar concentrator. Due to the higher temperatures gener- 7.55%, respectively. Just like the power output density, it is also
ated in systems 5/6, it generates the highest values compared to all seen that while varying the leg geometry increases the optimum
other systems. Also, system 1 records the least values, while system load requirement for maximum efficiencies in a rectangular leg
4 gives higher values compared to systems 2/3 combined. This tells TEG, multi-staging reduces this requirement. Furthermore, the crit-
that the use of performance enhancement strategies such as variable ical load requirement for the efficiencies in systems 2/3 to exceed
leg geometries and multi-staging reduces the thermodynamic stabi- that of system 1 is 1.2 Ω, while that for systems 5/6 to exceed
lity of the conventional single-stage rectangular leg TEG. system 4 is 0.9 Ω, thus implying that the efficiencies obtained
Finally, Fig. 6(g) shows that the exergy conversion index is max- from a variable leg geometry TEG will only surpass that of its rec-
imized at an optimum concentrated solar radiation intensity. This tangular leg counterpart beyond a specific critical load resistance.
new optimum value is different from that required for maximum Once again, multi-staging reduces this critical load requirement
exergy efficiency as earlier discussed in the previous paragraph, by 25%. Finally, comparing the optimum values recorded in this
except for system 1. This tells that there is an optimum solar radiation paragraph with that of the previous one shows that the optimum
at which maximum conversion of input exergy to useful power load requirements for maximum efficiencies are higher than that
output can be obtained from a solar TEG. Furthermore, due to the dif- for maximum power output densities. Numerically speaking, for
ferent optimum operation parameters of the various systems, this systems 1, 2/3, 4, and 5/6, the optimum load required for
optimum solar flux varies across the various systems studied. Typi- maximum power must be increased by 14.3%, 15.38%, 16.67%,
cally, the optimum solar flux for maximum exergy conversion in and 18.18%, respectively, before maximum efficiency can be
systems 1 and 2–6 are 15 and 10, respectively. The corresponding obtained. This further suggests that while the conventional system
maximum conversion indices obtained from systems 1, 2/3, 4, and 1 requires the least load resistance percentage increase for
5/6 are 1.06, 1.04, 2.04, and 2.02, respectively. First, this reveals maximum efficiency, systems 5/6 requires the highest.
that the adaption of multi-staging and tapered legs reduces the The irreversibilities and entropy generation densities associated
amount of solar flux and solar concentrator cost required to maximize with varying the load resistance are seen in Figs. 7(d) and 7(e),
exergy conversion in system 1 by 33.33% (one-third). However, respectively. Due to the large marginal differences in the values
compared to system 1, systems 2/3 reduce the exergy conversion generated by the various systems, we depict the variation of
index by 1.89%, while systems 5/6 reduce maximum exergy conver- system 1—as an example—to show the true behavior of these
sion of system 4 by 0.98%. These reductions are due to the higher parameters. Figure 7(f ) shows this variation. The need for load
exergy destruction recorded in the latter systems compared to the resistance optimization is seen in these plots, in that the least
former. However, the percentage reduction is quite low. Since our system irreversibilities and entropies are generated at the
focus is to maximize the exergy efficiencies, we select a concentrated optimum load resistance for maximum efficiencies as reported in
solar radiation of 20 suns as the optimum flux. the previous paragraph. Taking system 1 as an example, it can be
seen that the least irreversibilities are obtained when the load resis-
tance is set to 0.8 Ω. This coincides with the optimum load resis-
4.3 Effect of External Load Resistance. We cannot conclude tance for maximum efficiency in system 1 as discussed earlier.
this section without investigating the impact of the external load resis- These findings are in perfect agreement with the second law of ther-
tance to which the TEG is delivering its generated current. This is modynamics. That is, maximum exergy efficiency is obtained when
because the power output, and consequently the overall systems per- the system irreversibilities and entropies are minimized. The same
formance, is a function of the magnitude of the load resistance. In also applies to all other systems studied.
addition, the current generated by the TEG is also a function of the Finally, we wrap up the results obtained from this study by
load resistance. Thus, a parabolic relationship will ensue between recording the exergy conversion indices of the various systems as
the power output and load resistance. This suggests an optimum the load resistance is varied. Figure 7(g) shows this trend. It is
value. Hence, this information will be of importance to the TEG seen that the optimum loads required to maximize the exergy con-
designer and manufacturer. Using an optical concentration ratio of version indices coincide with those for maximum conversion effi-
20 suns, we vary the external load resistance from 0 to 2.3 Ω. This ciencies. This further shows the need for an optimization study,
range is selected in order to cover no load conditions, while surpassing since it does not just maximize the exergy efficiency but also max-
the internal electrical resistance offered by the thermoelectric legs. imizes the conversion of the input solar exergy to useful power
In line with our predictions, Figs. 7(a)–7(c) show that there exists output. The corresponding maximum exergy conversion indices
an optimum load resistance at which peak power output density and across systems 1, 2/3, 4, and 5/6 are 1.06, 1.04, 2.04, and 2.02,
energy/exergy efficiencies, respectively, are obtained from a loaded respectively, thus implying that varying the leg geometry of a
TEG. The first observation made is that the optimum load varies single/double-stage rectangular leg TEG reduces its exergy conver-
across the diverse systems. Typical values show that for systems 1, sion index, while increasing the number of stages increase it.
2/3, 4, and 5/6, the optimum loads for maximum power output den-
sities are 0.7 Ω, 1.3 Ω, 0.6 Ω, and 1.1 Ω, respectively. The corre-
sponding maximum power output densities are 7.77 × 105 W m−3,
8.67 × 105 W m−3, 4.66 × 105 W m−3, and 4.74 × 105 W m−3, 5 Conclusions
respectively. Hence, utilizing variable leg geometries increases the The comprehensive numerical optimization of a novel TTEG
optimum load requirement for maximum power in a conventional with variable area leg geometries was implemented using the

041303-12 / Vol. 144, APRIL 2022 Transactions of the ASME


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18
commercial software, ANSYS 2020 R2. The leg geometries consid- S = entropy generation, W K−1
ered were the traditional rectangular leg, optimum trapezoidal leg, T = temperature, K
and the recently introduced X-legs. The optimized parameters Ex = exergy
were the leg height and area, concentrated solar flux, and external
load resistance, thus encompassing the geometric, thermal, and
electrical performance of the device. These major conclusions are Greek Symbols
drawn from the study:
α = Seebeck coefficient, V K−1
• For the same amount of incident solar flux, the temperatures ɛ = emissivity
recorded across the hot and cold junctions of the tapered leg/ η = efficiency, %
rectangular leg and single/double-stage TEGs were all differ- μ = exergy conversion index
ent. This implies that future studies should avoid assuming ρ = electrical resistivity (Ω m)
the same temperatures across these various models. σ = Stefan–Boltzmann constant, W m−2 K−4
• The proposed two-stage TEG with tapered legs (trapezoidal τ = Thomson effect, transmissivity, V K−1
and X-legs) improved the exergetic efficiency of the base ϕ = electric field scalar potential, V
case, single-stage rectangular leg TEG, by 16.7%. Also, the ψ = concentrated solar radiation intensity, Suns
use of tapered leg TEGs, in single and multi-stage arrange-
ments, reduced the exergy conversion index of conventional
rectangular leg TEGs by 1.89% and 0.98%, respectively. Subscripts
• The results of the geometric optimization divulged that to
obtain maximum efficiencies, the initial leg height of the ther- a = ambient
moelectric legs, 1.62 mm, was increased by 517.28%, while c = cold
the initial leg area, 1.96 mm2, was decreased by 64.29%. d = destroyed
• The optimum leg height, area, and concentrated solar flux were en = energy
10 mm, 0.7 mm2, and 20 suns, respectively. However, the ex = exergy
optimum load resistances for systems 1, 2/3, 4, and 5/6 are f = inflow
0.8 Ω, 1.5 Ω, 0.7 Ω, and 1.3 Ω, respectively. These values cor- g = glass
respond to thermal/energy/exergy efficiencies of 7.05%/ gen = generation
6.03%/6.47%, 6.93%5.93%/6.36%, 8.08%/6.9%/7.41%, and h = hot
8.23%/7.03%/7.55%, respectively. rect-leg = rectangular leg geometry
• An increase in the leg height of the TEG increased its energy/ s = selective absorber surface, sun
exergy efficiencies while decreasing its power output per unit SB = Stefan–Boltzmann
volume. Thus, economic considerations must be made by the sky = sky
TEG manufacturer in the selection of the appropriate leg te = thermoelectric
height in order to achieve a high-power output at a reasonable X-leg = X-leg geometry
efficiency.
• X-leg TEGs are suitable for maximum power application (leg
height below 3 mm). For maximum efficiency application (leg References
height above 3 mm), either trapezoidal or X-leg TEGs can be [1] Zou, C., Zhao, Q., Zhang, G., and Xiong, B., 2016, “Energy Revolution: From a
used. Fossil Energy Era to a New Energy Era,” Nat. Gas Ind. B, 3(1), pp. 1–11.
• The attainment of maximum exergy efficiency in a solar TEG [2] Khan, M. Q., Malarmannan, S., and Manikandaraja, G., 2018, “Power Generation
From Waste Heat of Vehicle Exhaust Using Thermo Electric Generator: A
does not imply that all the input solar exergy is converted to Review,” IOP Conference Series: Materials Science and Engineering,
useful power output. Thus, the novel exergy conversion Kattankulathur, India, March, pp. 1–9.
index introduced fully captured the requirements for [3] Kotcher, J., Maibach, E., and Choi, W. T., 2019, “Fossil Fuels Are Harming Our
maximum exergy conversion in the device. Brains: Identifying Key Messages About the Health Effects of Air Pollution From
Fossil Fuels,” BMC Public Health, 19(1), pp. 1–12.
• Thinner legs were crucial to obtaining maximum overall per- [4] Moriarty, P., and Honnery, D., 2016, “Can Renewable Energy Power the
formance in a TEG, although this was associated with a Future?,” Energy Policy, 93, pp. 3–7.
decrease in the systems’ thermodynamic stability. [5] Shafiee, S., and Topal, E., 2009, “When Will Fossil Fuel Reserves Be
• Although the adoption of multi-stage configuration improved Diminished?,” Energy Policy, 37(1), pp. 181–189.
[6] Jaziri, N., Boughamoura, A., Müller, J., Mezghani, B., Tounsi, F., and Ismail, M.,
the single-stage system overall performance, it increased its 2019, “A Comprehensive Review of Thermoelectric Generators: Technologies
optimum concentrated solar flux requirement and ultimately and Common Applications,” Energy Rep., 24.
cost, while reducing the system thermodynamic stability. [7] Champier, D., 2017, “Thermoelectric Generators: A Review of Applications,”
• The power output densities and efficiencies obtained from a Energy Convers. Manage., 140, pp. 167–181.
[8] Lee, H., 2017, Thermoelectrics, 1st ed., John Wiley & Sons Ltd, West Sussex.
variable leg geometry TEG will only surpass that of its rectan- [9] Thimont, Y., and LeBlanc, S., 2019, “The Impact of Thermoelectric Leg
gular leg counterpart beyond a specific critical load resistance. Geometries on Thermal Resistance and Power Output,” J. Appl. Phys., 126(9),
p. 095101.
[10] Ali, H., Yilbas, B. S., and Sahin, A. Z., 2015, “Exergy Analysis of a
Thermoelectric Power Generator: Influence of Bi-Tapered Pin Geometry on
Nomenclature Device Characteristics,” Int. J. Exergy, 16(1), pp. 53–71.
h = heat transfer coefficient, W m−2 K−1 [11] Liu, H., Wang, S.-L., Yang, Y., Chen, W., and Wang, X., 2020, “Theoretical
k = thermal conductivity, W m−1 K−1 Analysis of Performance of Variable Cross-Section Thermoelectric Generators:
Effects of Shape Factor and Thermal Boundary Conditions,” Energy, 201,
v = wind speed, m s−1 p. 117660.
A = area, m2 [12] Shittu, S., Li, G., Tang, X., Zhao, X., Ma, X., and Badiei, A., 2020, “Analysis of
C = concentration ratio Thermoelectric Geometry in a Concentrated Photovoltaic-Thermoelectric Under
G = global solar radiation, W m−2 Varying Weather Conditions,” Energy, 202, pp. 1–13.
[13] Lamba, R., and Kaushik, S. C., 2017, “Thermodynamic Analysis of
I = current, A Thermoelectric Generator Including Influence of Thomson Effect and Leg
J = current density, A m−2 Geometry Configuration,” Energy Convers. Manage., 144, pp. 388–398.
K = thermal conductance, W K−1 [14] Ferreira-Teixeira, S., and Pereira, A. M., 2018, “Geometrical Optimization of a
Thermoelectric Device: Numerical Simulations,” Energy Convers. Manage.,
P = power output, W 169, pp. 217–227.
Q = heat flow, W [15] Ibeagwu, O. I., 2019, “Modelling and Comprehensive Analysis of TEGs With
R = internal electrical resistance, Ω Diverse Variable Leg Geometry,” Energy, 180, pp. 90–106.

Journal of Energy Resources Technology APRIL 2022, Vol. 144 / 041303-13


om http://asmedigitalcollection.asme.org/energyresources/article-pdf/144/4/041303/6729963/jert_144_4_041303.pdf?casa_token=JIgzEuB09zUAAAAA:9b7O5cxnPSzr333oqdqW3VK4fQTWuxZidck5LYzmiiHff8g8awiyqB7nQjz3_KzACpi1MC_l by National Institute of Technology- Surathkal user on 18
[16] Wang, R., Meng, Z., Luo, D., Yu, W., and Zhou, W., 2020, “A Comprehensive [35] Shittu, S., Li, G., Zhao, X., Ma, X., Akhlaghi, Y. G., and Fan, Y., 2020,
Study on X-Type Thermoelectric Generator Modules,” J. Electron. Mater., 49(7), “Comprehensive Study and Optimization of Concentrated Photovoltaic-
pp. 4343–4354. Thermoelectric Considering All Contact Resistances,” Energy Convers.
[17] Niu, Z., Yu, S., Diao, H., Li, Q., Jiao, K., Du, Q., Tian, H., and Shu, G., 2015, Manage., 205, p. 112422.
“Elucidating Modeling Aspects of Thermoelectric Generator,” Int. J. Heat Mass [36] Maduabuchi, C. C., Mgbemene, C. A., and Ibeagwu, O. I., 2020, “Thermally
Transfer, 85, pp. 12–32. Induced Delamination of PV-TEG: Implication of Leg’s Joule and Thomson
[18] Fabián-mijangos, A., Min, G., and Alvarez-quintana, J., 2017, “Enhanced Heating,” J. Electron. Mater., 49(11), pp. 6417–6427.
Performance Thermoelectric Module Having Asymmetrical Legs,” Energy [37] Maduabuchi, C., Ejenakevwe, K., Ndukwe, A., and Mgbemene, C., 2021, “High
Convers. Manage., 148, pp. 1372–1381. Performance Solar Thermoelectric Generator Using Asymmetrical Variable Leg
[19] Maduabuchi, C., Lamba, R., Njoku, H., Eke, M., and Mgbemene, C., 2021, Geometries,” E3S Web Conference, Rome, Italy, Vol. 239, p. 00005.
“Effects of Leg Geometry and Multistaging of Thermoelectric Modules on the [38] Moran, M. J., Shapiro, H. N., Boettner, D. D., and Bailey, M. B.,
Performance of a Photovoltaic-Thermoelectric System Using Different 2018, Fundamentals of Engineering Thermodynamics, 9th ed., Wiley,
Photovoltaic Cells,” Int. J. Energy Res., p. er.6925. Hoboken, NJ.
[20] Cheng, K., Qin, J., Jiang, Y., Lv, C., Zhang, S., and Bao, W., 2018, “Performance [39] Maduabuchi, C. C., and Mgbemene, C. A., 2020, “Numerical Study of a Phase
Assessment of Multi-Stage Thermoelectric Generators on Hypersonic Vehicles at Change Material Integrated Solar Thermoelectric Generator,” J. Electron.
a Large Temperature Difference,” Appl. Therm. Eng., 130, pp. 1598–1609. Mater., 49(10), pp. 5917–5936.
[21] Zhang, H., Xu, H., Chen, B., Dong, F., and Ni, M., 2017, “Two-Stage [40] Maduabuchi, C. C., and Mgbemene, C. A., 2020, “Numerical Analysis and
Thermoelectric Generators for Waste Heat Recovery From Solid Oxide Fuel Simulation of a Hybrid Concentrated Thermoelectric Module With Phase
Cells,” Energy, 132, pp. 280–288. Change Material,” 2020 Sustainable Engineering and Industrial Technology
[22] Cheng, K., Zhang, D., Qin, J., Zhang, S., and Bao, W., 2018, “Performance Conference, Nsukka, Enugu, July, pp. 1–3.
Evaluation and Comparison of Electricity Generation Systems Based on Single- [41] Lamba, R., Manikandan, S., and Kaushik, S. C., 2018, “Performance Analysis
and Two-Stage Thermoelectric Generator for Hypersonic Vehicles,” Acta and Optimization of Concentrating Solar Thermoelectric Generator,”
Astronaut., 151, pp. 15–21. J. Electron. Mater., 47(9), pp. 5310–5320.
[23] Lee, M.-Y., Seo, J., Lee, H., and Garud, K. S., 2020, “Power Generation, [42] Petela, R., 2003, “Exergy of Undiluted Thermal Radiation,” Sol. Energy, 74(6),
Efficiency and Thermal Stress of Thermoelectric Module With Leg Geometry, pp. 469–488.
Material, Segmentation and Two-Stage Arrangement,” Symmetry (Basel), [43] Maduabuchi, C., Ejenakevwe, K., Jacobs, I., and Ndukwe, A., 2020, “Analysis of
12(5), p. 786. a Two-Stage Variable Leg Geometry Solar Thermoelectric Generator,” 2nd
[24] Liu, Z., Zhu, S., Ge, Y., Shan, F., Zeng, L., and Liu, W., 2017, “Geometry African International Conference and Industrial Engineering and Operations
Optimization of Two-Stage Thermoelectric Generators Using Simplified Management, Harare, Zimbabwe, November, pp. 1–7.
Conjugate-Gradient Method,” Appl. Energy, 190, pp. 540–552. [44] Maduabuchi, C. C., Eke, M. N., and Mgbemene, C. A., 2021, “Solar Power
[25] Cheng, K., Qin, J., Jiang, Y., Zhang, S., and Bao, W., 2018, “Performance Generation Using a Two-Stage X-Leg Thermoelectric Generator With
Comparison of Single- and Multi-Stage Onboard Thermoelectric Generators High-Temperature Materials,” Int. J. Energy Res., 45(9), pp. 13163–13181.
and Stage Number Optimization at a Large Temperature Difference,” Appl. [45] Maduabuchi, C. C., Ejenakevwe, K. A., and Mgbemene, C. A., 2021,
Therm. Eng., 141, pp. 456–466. “Performance Optimization and Thermodynamic Analysis of Irreversibility in a
[26] Zhang, F., Cheng, L., Wu, M., Xu, X., Wang, P., and Liu, Z., 2020, “Performance Contemporary Solar Thermoelectric Generator,” Renew. Energy, 168,
Analysis of Two-Stage Thermoelectric Generator Model Based on Latin pp. 1189–1206.
Hypercube Sampling,” Energy Convers. Manage., 221, p. 113159. [46] Maduabuchi, C., Njoku, H., Eke, M., Mgbemene, C., Lamba, R., and Ibrahim,
[27] Kanimba, E., Pearson, M., Sharp, J., Stokes, D., Priya, S., and Tian, Z., 2017, “A J. S., 2021, “Overall Performance Optimisation of Tapered Leg Geometry
Modeling Comparison Between a Two-Stage and Three-Stage Cascaded Based Solar Thermoelectric Generators Under Isoflux Conditions,” J. Power
Thermoelectric Generator,” J. Power Sources, 365, pp. 266–272. Sources, 500, p. 229989.
[28] Sun, H., Ge, Y., Liu, W., and Liu, Z., 2019, “Geometric Optimization of [47] Ust, Y., Sahin, B., and Sogut, O. S., 2005, “Performance Analysis and
Two-Stage Thermoelectric Generator Using Genetic Algorithms and Optimization of an Irreversible Dual-Cycle Based on an Ecological Coefficient
Thermodynamic Analysis,” Energy, 171, pp. 37–48. of Performance Criterion,” Appl. Energy, 82(1), pp. 23–39.
[29] Guo, X., Zhang, H., Wang, J., Zhao, J., Wang, F., Miao, H., Yuan, J., and Hou, S., [48] Akkaya, A., Sahin, B., and Huseyinerdem, H., 2007, “Exergetic Performance
2020, “A New Hybrid System Composed of High-Temperature Proton Exchange Coefficient Analysis of a Simple Fuel Cell System,” Int. J. Hydrogen Energy,
Fuel Cell and Two-Stage Thermoelectric Generator With Thomson Effect: 32(17), pp. 4600–4609.
Energy and Exergy Analyses,” Energy, 195, p. 117000. [49] Ibeagwu, O. I., Eke, M. N., Maduabuchi, C. C., Mgbemene, C. A., and Aka, T. V.,
[30] Mgbemene, C. A., Njoku, H. O., and Agbo, C. O. A., 2018, “Investigation of 2020, “Particle Overlay Obstruction Modelling, Parametric and Output
Parametric Performance of the Hybrid 3D CPC/TEM System Due to Characteristics Evaluation of a Photovoltaic System,” Niger. Res. J. Eng.
Thermoelectric Irreversibilities,” Front. Energy Res., 6, pp. 1–11. Environ. Sci., 5, pp. 679–693.
[31] Erturun, U., Erermis, K., and Mossi, K., 2014, “Effect of Various Leg Geometries [50] Lee, H., 2010, Thermal Design Heat Sinks, Thermoelectrics, Heat Pipes,
on Thermo-Mechanical and Power Generation Performance of Thermoelectric Compact Heat Exchangers, and Solar Cells, 1st ed., John Wiley & Sons Ltd,
Devices,” Appl. Therm. Eng., 73(1), pp. 126–139. Hoboken, NJ.
[32] Maduabuchi, C., and Eke, M., 2021, “Solar Electricity Generation Using a [51] Shittu, S., Li, G., Zhao, X., Ma, X., Akhlaghi, Y. G., and Ayodele, E., 2019,
Photovoltaic-Thermoelectric System Operating in Nigeria Climate,” IOP “High Performance and Thermal Stress Analysis of a Segmented Annular
Conference Series: Earth and Environmental Science, University of Nigeria, Thermoelectric Generator,” Energy Convers. Manage., 184, pp. 180–193.
Nsukka, Enugu State, Nigeria, November 2020, Vol. 730, p. 012029. [52] Shittu, S., Li, G., Xuan, Q., Xiao, X., Zhao, X., Ma, X., and Akhlaghi,
[33] Xuan, X., Ng, K., Yap, C., and Chua, H., 2002, “The Maximum Temperature Y. G., 2020, “Transient and Non-Uniform Heat Flux Effect on Solar
Difference and Polar Characteristic of Two-Stage Thermoelectric Coolers,” Thermoelectric Generator With Phase Change Material,” Appl. Therm. Eng.,
Cryogenics (Guildf), 42(5), pp. 273–278. 173, pp. 1–15.
[34] Shittu, S., Li, G., Xuan, Q., Zhao, X., Ma, X., and Cui, Y., 2020, “Electrical and [53] Xiao, J., Yang, T., Li, P., Zhai, P., and Zhang, Q., 2012, “Thermal Design and
Mechanical Analysis of a Segmented Solar Thermoelectric Generator Under Management for Performance Optimization of Solar Thermoelectric
Non-Uniform Heat Flux,” Energy, 199, p. 117433. Generator,” Appl. Energy, 93, pp. 33–38.

041303-14 / Vol. 144, APRIL 2022 Transactions of the ASME

You might also like