You are on page 1of 20

Heat Transfer—Asian Research, 43 (5), 2014

Thermodynamic and Economic Optimization of Plate Fin Heat


Exchangers Using the Bees Algorithm

Salem Banooni, Hossein Zarea, and Maysam Molana


Department of Mechanical Engineering, Engineering Faculty, Shahid Chamran University,
Ahvaz, Iran

This study presents the successful application of the bees algorithm (BA) for
optimal design of a cross-flow plate fin heat exchanger by offset strip fins. The ε –
NTU method is used to approximate the heat exchanger effectiveness and pressure
drop. Two different objective functions including the minimization of total annual cost
(sum of investment and operational costs) and total number of entropy generation units
for certain heat duty required under given space constraints are considered as targets
of optimization separately. Based on the applications, seven design parameters (heat
exchanger length at hot and cold sides, fin height, fin frequency, fin thickness, fin-strip
length, and number of hot side layers) are selected as optimization variables. Two
examples from the literature are presented to illustrate the efficiency and accuracy of
the proposed algorithm. Results showed that the BA can detect an optimum
configuration with higher speed (short computational time) and accuracy compared
to the imperialist competitive algorithm (ICA) and the genetic algorithm (GA). ©
2013 Wiley Periodicals, Inc. Heat Trans Asian Res, 43(5): 427–446, 2014;
Published online 3 October 2013 in Wiley Online Library (wileyonlinelibrary.com/
journal/htj). DOI 10.1002/htj.21087

Key words: plate fin heat exchanger, optimization, bees algorithm, imperialist
competitive algorithm, genetic algorithm

1. Introduction

Different types of heat exchangers are applied for various industrial applications. One of the
important types is the compact heat exchanger (CHE). CHEs are described by a high heat transfer
surface area per unit volume (higher than 700 m2/m3) of the exchanger and can be either tube-fin type
or plate-fin (PFHE) type [1]. These heat exchangers are widely used in gas-to-gas heat exchange in
automobiles, cryogenics (refrigeration in very low temperature), aerospace and chemical, petroleum,
and petrochemical industries.

Extended surfaces or fins are components that decrease size and increase heat transfer and are
widely used in CHEs [2, 3]. Some of the more popular fin types are offset strip, louver, wavy, pin,
and perforated fins [4]. The boundary layer in offset-strip fins is discontinuous. Since the average
boundary layer thickness decreases significantly by discontinuity of boundary layer when offset-strip
fins are used, the convection heat transfer coefficient increases. Moreover, the fins with a rectangular
© 2013 Wiley Periodicals, Inc.
4271
cross section have a higher strength than do triangular fins [5]. Therefore, a rectangular offset-strip
fin, because of its excellent thermal efficiency, is used in this study.

However, offset-strip fins have high pressure drop. Therefore, the optimal design of a CHE
is always required to resolve the optimal conflict between the heat transfer rate and the power
consumption due to higher pressure drop. Hence, analysis based on the second law of thermodynamics
can be used for this purpose and the best way to evaluate this situation is by consideration of both
conflict factors [6–8]. The second-law-based entropy generation minimization (EGM) technique is
applied as a criterion for optimal analysis to appraise the thermal efficiency of each actual system that
owes its defects to fluid flow, heat transfer, mass transfer, and other transport processes. In a heat
exchanger, irreversibility is generated due to the limited temperature difference between the fluid
streams and the pressure drops along them. The entropy generation as a quantitative gauge of the
irreversibility is related to the heat transfer and fluid friction during the heat exchanger operation and
serves as a direct measure of the lost potential for work or the lost ability to transfer heat in the case
of a heat exchanger.

The number of entropy generation units (Ns) indicates this irreversibility associated with the
lost work or lost heat. Thus, minimizing the number of entropy generation units (EGU) means
minimizing the irreversibility of the system which increases the available part of the work or heat
transfer [7–10].

The design of a PFHE is a complex process based on a trial-and-error method in which


geometrical and operational parameters are selected to satisfy specified requirements such as outlet
temperature, heat duty, and pressure drop. Moreover, optimization based on the requested target
should always be taken into account. According to the literature, the customary objectives in heat
exchanger design are associated with minimizing investment and operational costs. Generally, a
higher flow velocity means a higher heat transfer coefficient and thus a lower heat transfer area and
consequently lower investment cost. Also, it should be noted that higher velocity leads to higher
pressure drop and thus higher power consumption and hence higher pumping cost. So, before doing
any optimal design, a suitable objective function is needed. In most cases, an agreement between the
investment cost and operational cost should be attained by the design variables. Therefore, minimizing
the total annual cost is discussed as the objective function [6].

Many works have been dedicated to the optimization of heat exchangers using traditional
mathematical techniques [11–15]. Recently, application of evolutionary algorithms has gained much
attention in the design of heat exchangers. In the early efforts, GA has been successfully used for
optimization of the shell-and-tube heat exchanger [16, 17] and obtaining heat transfer correlations for
CHEs [18, 19]. In the case of plate fin heat exchangers, a single-objective GA and PSO (particle
swarm optimization) have been used for optimizing plate fin heat exchangers [20–25] aiming at
minimization of a variety of objectives such as the ratio of the number of heat transfer units (NTU)
to the cold side pressure drop, total annual cost, total volume, total weight, and the number of entropy
generation units. Moreover, some works aimed at multi-objective optimization [26–29] used GA.

Recently, Yousefi and colleagues [30] used an imperialist competitive algorithm (ICA) to
optimize a PFHE considering minimization of total annual cost and total weight under given
constrained conditions. Comparing their result to the GA ones, they demonstrated that ICA achieves

428
shorter computational time and better results for their case. Also, they [31] employed an ICA to
optimize a cross-flow plate fin heat exchanger with the goal of minimizing the entropy generation
units (EGU). Their results showed that the performance of ICA is better than traditional GA.

Pham and colleagues [32] examined the first application of the bees algorithm for optimal
design of mechanical problems. They considered two standard samples, the design of welded beams
and helical spring. The purpose of that work was an examination of the bees algorithm performance
compared to other optimization algorithms. The results showed that the performance of the BA is
better than the others. Pham and Ghanbarzadeh presented the first application of the bees algorithm
for optimizing multi-objective problems. They studied the welded beams design problem using this
technique and found that it produces better results compared to other optimization algorithms [33].

In this study a plate fin heat exchanger is modeled. Total annual cost and the number of entropy
generation units are considered as two single-objective functions. As well, the sensitivity analysis of
changes in optimal value of the total annual cost with changes in design parameters is investigated
and results are presented. Finally, to show the efficiency of the proposed algorithm, besides the
objective functions mentioned in this article, the total heat transfer area and total pressure drop
objective functions, which depend on heat exchanger investment and operational costs, respectively,
are considered comparing them with the obtained results of Refs. 34 and 35.

Nomenclature

a: annual coefficient factor


A: heat exchanger surface area, m2
Aff: free-flow area, m2
C: heat capacity rate, W/K
Cp: specific heat, J/kg K
Cr: Cmin / Cmax
Dh: hydraulic diameter, m
EA: cost per unit area, $/m2
Ein: initial cost, $/year
Eop: operating cost, $/year
f: Fanning friction factor
f(x): objective function
g(x): constraint
G: mass flow velocity, kg/m2s
H: height of fin, m
h: convective heat transfer coefficient, W/m2K
j: Colburn factor
kel: electricity price, $/MWh
lf: Lance length of the fin, m
L: heat exchanger length, m
M: mass flow rate, kg/s
n: fin frequency, fin/m
n1: exponent of nonlinear increase with area increase
Na, Nb: number of fin layers for fluid a and b

429
Ns: number of entropy generation units (EGU)
NTU: number of transfer units
P: pressure, N/m2
Pr: Prandtl number
Q: heat duty, W
r: interest rate
R: specific gas constant, J/kg K
Rl: penalty parameter
Re: Reynolds number
s:
. fin spacing, m
S: rate of entropy generation, W/K
t: fin thickness, m
T: temperature, K
U: overall heat transfer coefficient, W/m2K
Vt: volumetric flow rate, m3/s
y: depreciation time

Greek Symbols

ε: effectiveness
η: efficiency of the fan or pump
μ: viscosity
ρ: density
τ: hours of operation
ΔP: pressure drop
ΔS: entropy difference, W/kg K

Subscripts

a, b: fluid a and b
i, j, r: variable number
1: inlet
2: outlet
max: maximum
min: minimum

2. Thermal Modeling

Figures 1 and 2 show a schematic of a plate fin heat exchanger and offset-strip fin with a
rectangular cross section, respectively. In order to simplify the analysis, the variation of thermophysi-
cal properties of fluids such as viscosity, Prandtl number, and specific heat with the temperature are
assumed negligible and both fluids are considered as an ideal gas. Other assumptions are as follows:

1. To minimize heat losses to the environment, the number of fin layers for the cold side
(Nb) is assumed to be one more than those of the hot side (Na).

2. Heat transfer coefficient and the area distribution are assumed to be constant and uniform.

430
Fig. 1. A schematic of a plate and fin heat Fig. 2. Typical rectangular offset-strip fin
exchanger (PFHE). core [25].

3. Heat exchanger works in steady-state conditions.

4. The thickness of the plates is assumed negligible. Therefore, thermal resistance and
longitudinal heat transfer of the walls are neglected.

5. For a gas-to-gas heat exchanger, the influence of fouling is neglected.


.
According to Bejan [36], the entropy generation rate (S) for two fluid streams can be expressed
in terms of temperature and pressure as

(1)

where Ta,2, Tb,2, Pa,2, and Pb,2 are the exit temperature and pressure of fluid a and fluid b, respectively,
which are found by considering the effectiveness of the heat exchanger as

(2)

So,

(3)

(4)

In the present work, because the outlet temperature of the fluids is not specified, the ε – NTU method
is employed in rating performance of the heat exchanger in the optimization process. For the
cross-flow heat exchanger with both fluids unmixed, effectiveness is given by Incropera and DeWitt
[37] as

(5)

431
where Cr = Cmin / Cmax and neglecting the thermal resistance of the walls and fouling factors, NTU is
calculated as follows:

(6)

(7)

Heat transfer coefficient is calculated from the Colbourn factor j:

(8)
In this formula G = m / Aff, where Aff is the free-flow cross-sectional area which is calculated
considering the geometrical details in Figs. 1 and 2.

(9)

(10)

Heat transfer area for both sides can be calculated similarly.

(11)

(12)

Then, total heat transfer area is given by

(13)
Heat transfer rate is calculated as

(14)
Frictional pressure drop in both sides is given by

(15)

(16)

There are many correlations for evaluation of the Colbourn factor j and Fanning factor f for
an offset strip fin. Equations (19) and (20) are the correlation presented by Manglik and Bergles [38]
which is used in this work.

(17)

432
(18)

where, by considering α = s / h, δ = t / lf, γ = t / s, and s = (1 / n − t) and h = (H − t), the hydraulic


diameter is

(19)

The Reynolds number is defined as

(20)

The above equations are valid for 120 < Re < 104, 0.134 < α < 0.997, 0.012 < δ < 0.048, and
0.041 < γ < 0.121. These equations correlate j and f factors from experimental data within +20%
accuracy in the laminar, transition, and turbulence flow regimes. Therefore, there is no need to describe
the flow regime for a specified operating condition and hence this is very useful in most practical
applications.

3. The Bees Algorithm

3.1 Bees in nature

A colony of honey bees can extend itself over long distances (more than 10 km) and in multiple
directions simultaneously to utilize a large number of food sources. A colony prospers by deploying
its foragers to good fields. In principle, flower patches with plentiful amounts of nectar or pollen that
can be collected with less effort should be visited by more bees, whereas patches with less nectar or
pollen should receive fewer bees.

The foraging process begins in a colony by scout bees being sent to search for promising
flower patches. Scout bees move randomly from one patch to another. During the harvesting season,
a colony continues its exploration, keeping a percentage of the population as scout bees.

When they return to the hive, those scout bees that found a patch which is rated above a certain
quality threshold (measured as a combination of some constituents, such as sugar content) deposit
their nectar or pollen and go to the “dance floor” to perform a dance known as the “waggle dance.”

This mysterious dance is essential for colony communication, and contains three pieces of
information regarding a flower patch: the direction in which it will be found, its distance from the
hive, and its quality rating (or fitness). This information helps the colony to send its bees to flower
patches precisely, without using guides or maps. Each individual’s knowledge of the outside
environment is gleaned solely from the waggle dance. This dance enables the colony to evaluate the
relative merit of different patches according to both the quality of the food they provide and the amount
of energy needed to harvest it. After waggle dancing on the dance floor, the dancer (i.e., the scout
bee) goes back to the flower patch with follower bees that were waiting inside the hive. More follower

433
bees are sent to more promising patches. This allows the colony to gather food quickly and efficiently
[39].

3.2 Proposed bees algorithm

The bees algorithm (BA) was developed by a group of researchers at the Manufacturing
Engineering Centre, Cardiff University [39]. This algorithm emulated the behavior of honey bees in
foraging for pollen and nectar to find the optimal solution for both continuous and combinatorial
problems. The algorithm required six parameters: the number of scout bees (n), number of selected
sites (m), number of top-ranking (elite) sites among the m selected sites (e), number of bees recruited
for each nonelite site (nsp), number of bees recruited for each elite site (nep), and neighborhood size
(ngh). The optimization process started with n scout bees randomly spread across the solution space.
Each scout bee was associated with a possible solution to the problem. The solutions were evaluated
and ranked in descending order of the fitness, and the best m sites were selected for a neighborhood
search.

In the neighborhood search procedure, more forager bees were sent in the neighborhood of
the elite (e) sites, and fewer bees around the nonelite (m-e) sites. According to this strategy, the
foraging effort was concentrated on the very best (i.e., elite) solutions. That is, nep bees were sent to
forage around the elite sites, while the area around the nonelite locations was exploited by nsp bees.
Within the given neighborhood area (i.e., flower patch size), some of the newly generated solutions
were expected to be better than those found by the scout bees. As shown in Fig. 3, only the best bee
is chosen to advertise its source after which the center of the neighborhood field is shifted to the
position of the best bee (i.e., from A to B).

In the global search procedure, the unselected scout bees (n-m) were used to explore at random
the solution space. This kind of search was to avoid bees being trapped at local optima. At the end of
each cycle, a new list of scout bees was formed, comprising the fittest solutions from each neighbor-
hood (neighborhood search results), and the new randomly generated solutions (global search results).
This list would be sorted in the next iteration and used for a new phase of optimization. The
combination of exploitative (neighborhood) and explorative (global) search would be able to capture

Fig. 3. Graphical explanation of the neighborhood search [40].

434
Fig. 4. Flowchart of the standard bees algorithm.

the best solution quickly and efficiently. These steps were repeated until the stopping criterion was
met [40]. The algorithm flowchart is shown in Fig. 4.

4. Objective Functions and Design Parameters

In the present study, the optimization targets are two single objective functions. The first is
from an economic point of view, the minimization of total annual cost, and the second objective has
been the minimization of the number of entropy generation units.

For the cost calculation, total annual cost is considered as the sum of investment cost and
operating cost. Investment cost is the annualized cost of the heat transfer area while operating cost
concerns the electricity cost for the compressors. The same approach for cost estimation was
considered by Refs. 6, 26, and 28:

(21)

(22)

(23)

where EA and n1 are cost per unit surface area and the exponent of nonlinear increase with area
increase, respectively. kel, τ, and η are the electricity price, hours of operation, and compressor
efficiency, respectively. Here, a is the annual coefficient factor that can be defined as

(24)

435
Table 1. Cost Coefficients of Heat Exchanger

where r and y represent interest rate and depreciation time, respectively. The parameters needed for
cost evaluation in this work are presented in Table 1. The second objective function, the number of
entropy generation units, is calculated as

(25)

Putting in all the relevant values, the above equation can be simplified and expressed as

(26)

5. Constraint Handling

A maximum of a function f is a minimum of –f. Thus, the general optimization problem may
be stated mathematically [33] as

(27)

(28)

where fi(x) are the l objective functions, X is the column vector of the k independent variables, and
Cj(X) are termed p equality constraints, and those of form hr(X) are q inequality constraints. In the
present work, the objectives are minimizing the total annual cost and the number of entropy generation
units. Moreover, to take into account the effect of constraints violation during the optimization
process, an arbitrarily large value (known as the penalty function) is also added in the objective
function [20, 26]. So, finally the objective function for the present work is represented as

Minimize f(x) = f(x) + ∑R1(gj(x))2 (29)


j=1

436
where R1 is the penalty parameter having a large value (say, 500). The term Σm
j=1R1(gj(x)) takes into
2

account the effect of constraints violation.

For all values of the penalty parameter R1, the bees algorithm is not sensitive to the penalty
function. So the right-hand term in the above equation is replaced with a large constant number (static
penalty function) similar to scheme 1 from Ref. 35.

6. A Case Study

To demonstrate the application of the mentioned optimization algorithm, similar to the


performed work of Yousefi and colleagues [30], this application example is taken from the work of
Shah and Sekulic [41]. A gas-to-air single-pass cross-flow heat exchanger having a heat duty of 1069.8
kW is needed to be designed and optimized for minimum total annual cost. Maximum of the exchanger
is limited to 1 × 1 × 1.5 m, gas and air inlet temperatures are 900 and 200 K, respectively, and the gas
and air mass flow rates are 1.66 and 2.00 kg/s, respectively. Pressure drops of the hot and cold sides
are set to be limited to 9.50 and 8.00 kPa, respectively. The gas and air inlet pressures are 160 and
200 kPa absolute. The offset-strip fin surface is used on the gas and air sides. The plate thickness is
set at 0.5 mm and is not an optimization variable. Operating conditions and the cost function constant
values needed for cost evaluation are listed in Table 2.

In this study, seven parameters—hot flow length (La), cold flow length (Lh), number of hot
side layers (Na), fin frequency (n), fin thickness (t), fin height (H), and lance length of fin (lf)—are
considered as optimization variables. All variables except the number of hot side layers are continu-
ous. The variation ranges of the variables are shown in Table 3. Additional inequality constraints are
set to guarantee that the no-flow length, pressure drops, and α, δ, γ at both sides maintain their
prescribed ranges. Moreover, another constraint is implementing to ensure that a minimum required
heat transfer is achieved.

7. Results and Discussion


7.1 Minimum total annual cost of the PFHE (case study A)

For the determined allowable pressure drop and heat duty requirement, the optimization
problem is detecting the design variable that minimizes the total annual cost of the PFHE. BA
parameters such as number of selected sites (m), number of top-ranking (elite) sites among the m

Table 2. Operating Parameters Selected for the Case Study

437
Table 3. Variation Range of Design Parameters

selected sites (e), number of bees recruited for each nonelite site (nsp), and number of bees recruited
for each elite site (nep) are set to 12, 9, 15, and 20, respectively. The expected results for the minimum
total annual cost are seen in Fig 5. Table 4 shows the optimum values of design parameters based on
minimization of the total annual cost. By comparing results of this study and the results of Yousefi
and colleagues [30], it can be concluded that the hot and cold flow length and the number of hot side
fin layers are decreased, the fin height and the fin frequency are increased, and both the free-flow
cross-sectional area of hot and cold flow are decreased 6.42% and 8.49%, respectively.

Therefore, the mass flow velocities of both the hot and cold sides are increased. On the other
hand, by increasing δ (due to reduced lance length of fin) and reducing α (due to increased fin height),
the friction factor and Colburn factor are increased. Therefore, a combination of these factors increases
the pressure drops on both hot and cold streams, 6.1% and 10.6%, respectively. Thus, the investment
cost and the heat transfer coefficient for both hot and cold streams are increased 7.3% and 9.3%,
respectively, compared to the results of Yousefi and colleagues [30].

It should be noted that with reducing the hot flow length, the free-flow area of hot flow remains
constant but the free-flow area of cold flow is reduced. So, the mass flow velocity and pressure drop
of the cold stream increased, although the pressure drop of the hot flow due to constant mass flow

Fig. 5. Convergence process of the total annual cost as objective function.

438
Table 4. Results of the ICA and BA for Minimum Total Annual Cost

velocity and a reduction of hot flow length is reduced. With a similar trend to reduce the cold flow
length, the pressure drop of the cold flow decreased and the pressure drop of the hot flow was
enhanced.

In this study, changes of the total pressure drop heat exchanger with variation of the hot and
cold flow length are almost identical. By reducing the hot and cold flow, and fin frequency, the fin
height total heat transfer surface area is decreased and hence the initial investment cost is reduced.
But because the initial investment cost includes a greater share of the total annual cost, the effect of
increasing pressure drops on the total annual cost is low. According to the mentioned reasons, in this
study, the total annual cost as an objective function decreased about 4.7% compared to the reference.
Moreover, the algorithm used in this study converged to the optimal value in less time. Thus, we
concluded that the used algorithm has higher accuracy and speed of convergence to the optimal value
of the total annual cost function, when compared to the imperialist competitive algorithm.

7.2 Sensitivity analysis

In this section, by considering the operating conditions and constant values given in Table 2,
the effect of a number of design parameters on the total annual cost objective function is evaluated.
In any section, the values of all parameters except the one that has been selected for the investigation
are kept constant at their optimum values. By changing the value of the selected parameter, the
sensitivity of any objective function can be investigated.

7.2.1 Effect of the fin height

The effect of the fin height on the total annual cost is shown in Fig. 6. By increasing the fin
height, both the free-flow area and the total heat transfer surface area are increased, and also by
reducing the value of α, the friction coefficient is decreased. Due to the decrease in mass flow velocity
from the increase in flow cross section, and also the decrease in the friction factor, pressure drop is
reduced which leads to lower values of the pumping power and the operational cost. As mentioned

439
Fig. 6. Effect of the fin height on the total annual cost.

earlier, any increase in the fin height increases the total heat transfer area and therefore the initial
investment cost of the PFHE. Thus, increasing the fin height has two conflicting effects on the total
annual cost of the system. Therefore, as is clear in Fig. 6, the initial variation of total annual cost with
respect to the fin height is descending, which shows that with increasing the fin height, a reduction
in the operational cost (pumping power) shows more of an increase in the initial investment cost. But
in the end, the variation in increased operating cost is similar to a reduction in investment cost, and
therefore the total annual cost almost remains constant.

7.2.2 Effect of the fin frequency

Figure 7 shows the variation of the total annual cost based on change in the fin frequency. The
increase in the fin frequency leads to the reduction in the free-flow area and the increase in the total

Fig. 7. Effect of the fin frequency on the total annual cost.

440
heat transfer area. Due to the reduction in the free-flow area, the mass flow velocity, pressure drop,
and consequently operational costs are increased. Also, due to the increase in total heat transfer area,
the initial investment cost is increased. Thus, the total annual cost is increased.

7.2.3 Effect of the number of fin layers

From the mentioned equations it can be deduced that each increase in the number of fin layers
leads to an enhancement of the total heat transfer surface and the free-flow area. Due to the increase
in free-flow area, the values of pressure drops are reduced, which in turn leads to a lower operational
cost. On the other hand, due to the increase in total heat transfer area, the initial investment cost is
also increased. Thus, increasing the number of fin layers has two conflicting effects on the total annual
cost. As can be seen in Fig. 8, the first variation of the total annual cost to the number of fin layers is
descending, which shows that with increasing the number of fin layers the reduction in the operational
cost (pumping power) is more than the increase in the initial investment cost. But after approximately
Nα = 65, the variation of the total annual cost to the number of fin layers is ascending and therefore,
by increasing the number of fin layers, the reduction in the operational cost is less than the increase
in the initial investment cost.

7.2.4 Effect of the lance length of fin

The increase in the fin offset length causes a decrease in δ, and thus the friction factor is
decreased. Due to a reduction in the friction factor, the total pressure drop and the operational cost
are decreased. In this condition, the heat transfer surface area remained constant and therefore had no
effect on the initial investment cost. Thus, according to Fig. 9, by increasing the fin-offset length, the
reduction process in the total annual cost and operational cost are almost the same.

Fig. 8. Effect of the number of fin layers on the total annual cost.

441
Fig. 9. Effect of the lance length of fin on the total annual cost.

7.3 Minimum number of entropy generation units of the PFHE (case study B)

This case is similar to case study A except that the maximum dimensions of the heat exchanger
are limited to 1 × 1 × 1 m and the heat exchanger needs to be designed and optimized for a minimum
number of entropy generation units. The case study considered is for comparison with Yousefi and
colleagues [31]. Table 5 shows the optimum values of the design parameters based on minimization
of the number of entropy generation units.

As is clear, despite the reduction in the fin-offset length (increase in the fin frequency) and
number of fin layers, by increasing the fin height, the free-flow area is increased. Therefore, the mass
flow velocity and pressure drop are reduced in both hot and cold streams, 39% and 31% versus the
results of ICA and 82% and 11% versus the results of GA, respectively. Hence, due to the reduction
in the pressure drop, hot and cold flow and also efficiency increase; the reduced number of entropy

Table 5. Comparison Results of the ICA, GA, and BA for Minimum


Number of Entropy Generation Units

442
Fig. 10. Convergence process of the number of entropy generation units as objective function.

generation units as an objective function in this study versus the results of ICA and GA were 2.5%
and 5.2%, respectively. Figure 10 shows the convergence of the number of entropy generation units
as an objective function.

7.4 A comparison among the proposed algorithm, ICA, GA, HIS, and GAHPSO

To demonstrate the effectiveness of the proposed algorithm, a comparison between this


algorithm and the algorithms of ICA and GA is discussed in terms of computational time and accuracy
convergence to the optimal value [30]. In addition to the objective functions provided in this article,
the objective functions of the total pressure drop and total heat transfer area, which are related to
operational cost and investment cost of the heat exchanger, respectively, have also been studied to
compare the genetic algorithm hybrid with the particle swarm optimization (GAHPSO) and an
improved harmony search algorithm with a static penalty parameter (IHS scheme 1) [34, 35]. Results
are shown in Table 6.

Table 6. Comparison Results of the ICA, GA, GAHPSO, HIS, and BA

443
Similar to ICA, GA, HIS, and GAHPSO, this algorithm’s population size and number of
iterations are set to 100 and 200, respectively. This algorithm is programmed in MATLAB® and run
on an Intel®CoreTMi5 CPU. The mentioned CPU time is an average of 10 executions of the computer
program. As is seen in case study A, the results of the BA were 2.7% and 10% better than those of
the ICA and GA, respectively. In case study B, the results of the BA were 2.5% and 5.2% better than
those of the ICA and GA, respectively. The same trend exists in case study C where the results of the
BA were similar to GAHPSO and IHS but were 25% better than those of the GA. Also, in case study
D, the results of the BA were 5.7%, 5.4% and 24.2% better than those of the GAHPSO, HIS, and GA,
respectively. In all three cases (A, C, D), the BA computational time was less than ICA, GA, HIS,
and GAHPSO. Also in all case studies, the results obtained by BA had higher accuracy. Therefore,
the BA has higher speed and accuracy and converges to the optimum value.

8. Conclusion

This study presents the successful application of a new algorithm for the optimal design of
plate fin heat exchangers. This algorithm is used in most thermal engineering problems that consist
of several discrete and continuous variables and a large amount of discontinuity in the objective
function. Depending on the applications, different design variables are optimized for minimum total
annual cost and number of entropy generation units under a given set of constraints. The constraints
are handled by adding a fixed penalty function to the fitness function. Two case studies from the
literature are selected for examination of the performance and accuracy of this new method. The results
reveal that the BA can find optimum configuration in less computational time under the same
population size and iterations. The design procedure for the PFHEs presented in this study by using
the BA can be applied to the other types of heat exchangers such as shell-and-tube heat exchangers.
Moreover, other types of fins such as plain, perforated, wavy, and louvered fins can be applied on
both the cold and hot sides of the heat exchanger rather than offset-strip fins which are applied on
both sides in the present work. The results can be used for designers to start with or to select an initial
guess.

Literature Cited

1. Kays WM, London AL. Compact heat exchangers, 3rd ed. McGraw-Hill; 1984.
2. Bergles AE. Techniques to enhance heat transfer. In: Rohsenow WM, Hartnett JP, Cho YI
(editors). Handbook of heat transfer, 3rd ed. McGraw–Hill; 1988. Chapter 11.
3. Manglik RM. Heat transfer enhancement. In: Bejan A, Kraus AD (editors). Heat transfer
handbook. Wiley; 2003. Chapter 14.
4. Kim B, Sohn B. An experimental study of flow boiling in a rectangular channel with offset
strip fins. Int J Heat Fluid Flow 2006;27:514–521.
5. Lihua G, Jiangping C, Feng Q, Zhijiu C. Geometrical optimization and mould wear effect on
HPD type steel offset strip fin performance. Energy Converse Manage 2007;48:2473–2480.
6. Xie GN, Sunden B, Wang QW. Optimization of compact heat exchangers by a genetic
algorithm. Appl Therm Eng 2008;28:895–906.
7. Mishra M, Das PK, Sarangi S. Second law based optimisation of cross-flow plate-fin heat
exchanger design using genetic algorithm. Appl Therm Eng 2009;29:2983–2989.
8. Sahiti N, Krasniqi F, Fejzullahu X, Bunjaku J, Muriqi A. Entropy generation minimization of
a double-pipe pin fin heat exchanger. Appl Therm Eng 2008;28:2337–2344.

444
9. Ko TH. Analysis of optimal Reynolds number for developing laminar forced convection in
double sine ducts based on entropy generation minimization principle. Energy Convers
Manage 2006;47:655–670.
10. Maximov JT. Thermodynamic optimization of mechanical systems with dissipative processes.
Int J Mech Sci 2006;48:62–74.
11. Reneaume JM, Niclout N. MINLP optimization of plate fin heat exchangers. Chem Biochem
Eng Q 2003;17:65–76.
12. Reneaume JM, Niclout N. Plate fin heat exchanger design using simulated annealing. Comput
Aided Chem Eng 2001;9:481–486.
13. Wang L, Sundén B. Design methodology for multi-stream plate-fin heat exchangers in heat
exchanger networks. Heat Transf Eng 2001;22:3–11.
14. Reneaume JM, Pingaud H, Niclout N. Optimization of plate fin heat exchangers: a continuous
formulation. Chem Eng Resour Des 2000;78:849–859.
15. Muralikrishna K, Shenoy UV. Heat exchanger design targets for minimum area and cost. Chem
Eng Resour Des 2000;78:161–167.
16. Selbas R, Kizilkan O, Reppich M. A new design approach for shell-and-tube heat exchangers
using genetic algorithms from economic point of view. Chem Eng Process 2006;45:268–275.
17. Tayal MC, Fu Y, Diwekar UM. Optimal design of heat exchangers: a genetic algorithm
framework. Ind Eng Chem Resour 1999;38:456–467.
18. Pacheco-Vega A, Sen M, Yang KT. Simultaneous determination of in- and over-tube heat
transfer correlations in heat exchangers by global regression. Int J Heat Mass Transf
2003;46:1029–1040.
19. Pacheco-Vega A, Sen M, Yang KT, Mclain RL. Genetic algorithms-based predictions of
fin-tube heat exchanger performance. Heat Transfer Proceeding of 11th IHTC, 1998, (6), p
137–145.
20. Peng H, Ling X. Optimal design approach for the plate-fin heat exchangers using neural
networks cooperated with genetic algorithms. Appl Therm Eng 2008;28(5–6):642–650.
21. Xie GN, Wang QW. Geometrical optimization design of plate-fin heat exchanger using genetic
algorithm. ZhongguoDianjiGongchengXuebao. Proc Chinese Soc Electr Eng 2006;26(7):53–
57.
22. Ozkol I, Komurgoz G. Determination of the optimum geometry of the heat exchanger body
via a genetic algorithm. Num Heat Transf Part A Appl 2005;48:283–296.
23. Mishra M, Das PK, Sarangi S. Optimum design of cross-flow plate-fin heat exchangers through
genetic algorithm. Int J Heat Exchangers 2004;5:379–401.
24. Peng H, Ling X, Wu E. An improved particle swarm algorithm for optimal design of plate-fin
heat exchangers. Ind Eng Chem Resour 2010;49:6144–6149.
25. Rao RV, Patel VK. Thermodynamic optimization of cross flow plate-fin heat exchanger using
a particle swarm optimization algorithm. Int J Therm Sci 2010;49:1712–1721.
26. Ahmadi P, Hajabdollahi H, Dincer I. Cost and entropy generation minimization of a cross-flow
plate fin heat exchanger using multi-objective genetic algorithm. J Heat Transf
2011;133:021801–021810.
27. Sanaye S, Hajabdollahi H. Thermal-economic multi-objective optimization of plate fin heat
exchanger using genetic algorithm. Appl Energy 2010;87:1893–1902.
28. Rao RV, Patel VK. Multi-objective optimization of heat exchangers using a modified teach-
ing-learning based optimization algorithm. Appl Math Model 2012;1–16.
29. Najafi H, Najafi B. Multi-objective optimization of a plate and frame heat exchanger via genetic
algorithm. Int J Heat Mass Transf 2010;46:639–647.
30. Yousefi M, Darus AN, Mohammadi H. An imperialist competitive algorithm for optimal
design of plate-fin heat exchangers. Int J Heat Mass Transf 2012;55:3178–3185.

445
31. Yousefi M, Darus AN, Mohammadi H. Second law based optimization of a plate fin heat
exchanger using imperialist competitive algorithm. Int J Phys Sci 2011;6:4749–4759.
32. Pham DT, Ghanbarzadeh A, Otri S. Optimal design of mechanical components using the bees
algorithm. Proc Inst Mech Eng Part C J Mech Eng Sci 2009;28:101–114.
33. Pham DT, Ghanbarzadeh A. Multi-objective optimisation using the bees algorithm. 3rd
International Virtual Conference on Intelligent Production Machines and Systems (IPROMS
2007), Whittles, Dunbeath, Scotland, 242, p 111–116.
34. Yousefi M, Enayatifar R, Darus AN. Optimal design of plate-fin heat exchangers by a hybrid
evolutionary algorithm. Int Commun Heat Mass Transf 2012;39:258–263.
35. Yousefi M, Enayatifar R, Darus AN, Abdullah AH. Optimization of plate-fin heat exchangers
by an improved harmony search algorithm. Appl Therm Eng 2012;50:877–885.
36. Bejan A. The concept of irreversibility in heat exchanger design: counter flow heat exchangers
for gas-to-gas applications. ASME J Heat Transf 1977;99:374–380.
37. Incropera FP, DeWitt DP. Fundamentals of heat and mass transfer. Wiley; 1998.
38. Manglik RM, Bergles AE. Heat transfer and pressure drop correlations for the rectangular
offset strip fin compact heat exchanger. Exp Therm Fluid Sci 1995;10:171–180.
39. Pham DT, Ghanbarzadeh A, Koc E, Otri S, Rahim S, Zaidi M. The bees algorithm. A novel
tool for complex optimization problems. Proceedings of the 2nd International Virtual Confer-
ence on Intelligent Production Machines and Systems, 2006, Elsevier. p 454–459.
40. Otri S. Improving the bees algorithm for complex optimization. PhD thesis, Cardiff University,
2011.
41. Shah RK, Sekulic DP. Fundamentals of heat exchanger design. Wiley; 2003.

"F F F"

446

You might also like