Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

Journal Pre-proofs

Numerical investigation of thermal and hydraulic performance of shell and


plate heat exchanger

Ali Abbas, Howard Lee, Akash Sengupta, Chi-Chuan Wang

PII: S1359-4311(19)34750-7
DOI: https://doi.org/10.1016/j.applthermaleng.2019.114705
Reference: ATE 114705

To appear in: Applied Thermal Engineering

Received Date: 10 July 2019


Revised Date: 9 November 2019
Accepted Date: 18 November 2019

Please cite this article as: A. Abbas, H. Lee, A. Sengupta, C-C. Wang, Numerical investigation of thermal and
hydraulic performance of shell and plate heat exchanger, Applied Thermal Engineering (2019), doi: https://
doi.org/10.1016/j.applthermaleng.2019.114705

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Ltd.


Numerical investigation of thermal and hydraulic performance of shell and plate heat
exchanger

Ali Abbas, Howard Lee, Akash Sengupta, Chi-Chuan Wang*


Department of Mechanical Engineering, National Chiao Tung University, Hsinchu,
Taiwan
*
Corresponding author: ccwang@nctu.edu.tw (C.-C. Wang).

Abstract
Detailed single-phase performance of the shell and plate heat exchangers is presented in this study
with corresponding chevron angles (β) of 45 and 75, respectively. Unlike those conventional
plate heat exchanger, the thermofluids characteristics perform quite differently between shell and
inner channel. For a larger chevron angle (75 ), the friction factor is comparable amid shell and
inner channel. Yet the friction factor in the shell channel is much lower than that in inner channel
at a smaller chevron angle (45). The shell channel contains more unidirectional flow pattern while
some flow re-circulation may occur at the entrance/exit port of the inner channel. The flow pattern
and temperature distribution in the shell channel are generally more uniform than those in inner
channel. The shell channel with β = 45° shows a better overall performance for its much lower
friction factor when the Reynolds number is low. The heat transfer performance for shell channel
exceeds the inner channel appreciably when the chevron angle is high (75). The heat transfer
performance depends on the corrugation aspect ratio, and an optimum ratio is derived. For the
inner channel, the optimum value is around 0.75 while it is 0.53 for the shell channel.

Keywords: Shell and plate heat exchanger; Chevron angle; heat transfer coefficient; friction factor.

1. Introduction

High-efficiency thermal energy conversion is especially needed for low-temperature

difference applications such as Ocean Thermal Energy Conversion (OTEC) plant where the

available temperature difference is as low as 15-25 C [1, 2]. Hence, ingenious equipment to

harvest the thermal energy is very imperative for applications with the low-temperature difference.

Regarding the energy harvesting and conversion equipment available, the shell-and-tube heat

exchangers (STHXs) are one of the mostly used for more than a century [3]. Despite STHXs offer

many superior characteristics such as robustness, reliability, and tolerable severe operating
conditions, their comparatively large size, less scalability, and relatively low thermal performance

limit their applicability in low temperature-difference energy harvesting. In this regard, an

alternative to STHX is the so-called plate heat exchangers (PHEs) which offer much higher heat

transfer performance and better scalability. Hence, PHEs become more popular in the past few

decades especially in some industrial applications such as dairy, food processing, paper/pulp,

heating, ventilating, and chemical processing [4]. For PHEs, some review papers [5, 6] had

summarized the modern technological advances and Zhang et al. [7] reviewed the heat transfer

enhancement techniques upon the plate heat exchangers via surface modification and nanofluids.

Based on their thorough reviews, the chevron angle (β) of the PHEs is the most influential

geometrical parameter for the single phase heat transfer applications. Yet the degree of importance

of β in the two-phase flow may not be as crucial as in single-phase applications but it still plays a

significant role.

There were numerous studies associated with the thermal and hydraulic performances of plate

heat exchangers. For example, Gulangyu et al. [8] experimentally investigated the thermal and

hydraulic characteristics of PHEs and reported a marked improvement of heat transfer

performance with decreasing plate size. Yang et al. [9] studied single-phase heat transfer

performance using ethylene glycol water mixture in both frame-and-plate (FPHE) and brazed plate

heat exchangers (BPHE) with chevron plates having U-type flow arrangement and developed a

correlation to predict their data. Yoon et al. [10] developed a numerical model based on flow

network approach to evaluate the performance of a plate heat exchanger (PHE). Turk et al. [11]

performed experiments regarding the thermal and hydraulic performance of gasket plate heat

exchangers (GPHE) using two analyzing methods (classical correlation and artificial neural

networks (ANNs)). They concluded that ANNs are easier to predict the performance of mixed
plate with more reliable predictions. Zhicheng et al. [12] focused on the shape optimization and

heat transfer analysis of a welded plate heat exchanger, and the effects of long axis, short axis, and

plate spacing were examined with the aid of CFD (Computational Fluid Dynamics). The influences

of various factors on the heat transfer characteristics were optimized using the grey correlation

theory. Song et al. [13] inspected the effects of both convex and concave vortex generators (VG)

in the plate heat exchangers. The results showed that the concave curved delta winglet VG is more

beneficial as far as heat transfer enhancement is concerned, while the convexly curved delta

winglet VG is disadvantageous to heat transfer enhancement. Jamzad et al. [14] used graphite as

the constructional material for plate heat exchanger owing to some favorable properties of graphite

(high thermal conductivity, light weight, and corrosion resistivity). When compared with the metal

alloy heat exchanger, the graphite plate heat exchanger shows identical thermal performance and

a 26% higher pressure drop. Elias et al. [15] conducted experiments to analyze the heat transfer

performance and pressure drop characteristics in a plate heat exchanger (PHE) having 30° and 60°

chevron angles with water-based Al2O3 nanofluid at the volume concentrations ranging from 0 to

0.5% vol. It was found that the addition of the nano-particles in working fluids enhances the heat

transfer rate. Kim et al. [16] performed experiments to inspect the heat transfer and pressure drop

characteristics on brazed plate heat exchanger (BPHE) having different geometries. Khanet al. [17]

studied the pressure drop characteristics of a gasket commercial plate heat exchanger for single-

phase. Their experimental results showed that mixed plate configuration could be a probable

choice in optimizing the plate heat exchanger design as far as performance is concerned. Ozkaya

et al. [18] performed CFD analysis to inspect the thermal-hydraulic characteristics of commercial

gasket plate heat exchangers with β = 30 and the numerical simulations showed good agreement

with the experimental results.


The foregoing studies indicated that numerous publications regarding PHEs from various

perspective efforts. Despite intensive efforts had been carried out, the conventional PHEs still

suffers some inherited drawbacks for its configuration (rectangular plate). The configuration

cannot withstand higher operating pressures and high temperatures, and is prone to leakage for

gasket type plate heat exchanger. Hence, in an endeavor to tackle the major shortcomings of PHEs,

the shell and plate heat exchangers (SAPHXs) evolved from plate heat exchanger and shell-and-

tube heat exchanger is developed as shown in Fig. 1. The SAPHXs contain a round shell which

encompassing numerous welded plates. The design is capable to withstand very high temperatures

and pressures in both shell and plate channel side, and the plate channel contains superior heat

transfer performance from its predecessor (PHE) that makes low-temperature thermal energy

harvesting feasible. Yet as depicted in Fig. 1, the plate bundle can be easily removed for cleaning

purpose (like PHE). Even though SAPHX already inherits outstanding features from its

predecessors and had been in existence for several decades, relevant researches on this topic is

very scarce. The only available studies relevant to SAPHX are Lim et al. [22], Nakaoka and Uehara

[23], Freire et al., [19], and Arsenyeva et al. [20]. The studies by Freire et al., [19] or Arsenyeva

et al. [20] just mentioned introductory material with regard to SAPHX. Lim et al. [21] investigated

the condensation heat transfer characteristics of R-245fa in SAPHX for high temperature heat

pumps. The condensation heat transfer coefficient and frictional pressure drop displayed an

incremental tendency with increasing mean vapor quality and mass flux and with decreasing

saturation temperature. Nakaoka and Uehara [2] conducted experiments for investigating the shell

and plate condenser for OTEC (Ocean thermal energy conversion) and proposed some empirical

correlations beneficial for predicting the average condensation heat transfer coefficient and the
water-side heat transfer coefficient. However, the plates used by Nakaoka and Tsutomu [2] were

gasket based rectangular plate which is unable to sustain high pressure applications.

To sum up, there were no available thermofluids characteristics about the SAPHX yet. As

depicted earlier, though SAPHX had existed for several decades and contains very superior

features, the heat exchangers were not popularly used. One of the specific reasons is lacking of

domain knowledge about the thermal characteristics of this kind heat exchanger and the

performance difference between SAPHX and the well-known PHE. Hence, the purpose of this

study is to bridge the gap by offering more detailed fluid flow and heat transfer performance for

the SAPHX through detailed numerical simulations. The provided details could elaborate thermal

energy harvesting having low-temperature difference applications like OTEC. Also, further

comparison regarding its superior thermal energy conversion and energy saving potential in

comparison with PHE is investigated. The proposed novel optimum SAPHX design not only offers

superior thermal energy conversion effectiveness in the range of 10~30% when compared to the

PHE but also features outstanding energy saving in pumping power. In fact, up to 64% energy-

saving in pumping power is attainable with the present design. In essence, the proposed design not

only contains superior thermofluids characteristics but also offers significant improvement in

thermal energy conversion with pronounced energy saving in the required pumping power. The

proposed design shows far more prominent feature when compared with the PHE as far as low-

temperature thermal energy conversion is concerned.


Fig. 1. Schematic of the shell and plate heat exchanger and the channel plate bundle.

(a) chevron angle (β) (b) inner channel structure

(a) Schematic of the plate and the inner channel.

(a) conventional sinusoidal corrugation (b) corrugation parameters

(b) Schematic of corrugated configuration.

Fig. 2. Diagram of the shell and plate geometry and the corrugation configuration.
2. Simulations and Data Reduction

The detailed schematic of present SAPHX is illustrated in [21]. The schematic of a single plate

and the single inner channel is illustrated in Fig. 2(a). In this study, the channels containing chevron

angles of 45 and the 75 are examined separately. Fig. 2(b) shows the conventional sinusoidal

configuration the plate configuration.

The computational domain of the inner channel and shell side channel, and their boundary are

shown in Figs. 3 (a) and 3(b) respectively. Half of the model domain was incorporated with

symmetry boundary condition at the fluid surfaces to reduce the mesh elements number. The inlet

and outlet ports were extended longitudinally four times of the port diameter to avoid any probable

flow reversal. The extended inlet and outlet ports were given a zero conductivity value. The

simulations are fulfilled with the commercial Autodesk CFD and Ansys fluent software to perform

more than 90 different cases, the software are based on the finite volume method. It was found that

results obtained from the two software are identical when the solution is mesh-independent. Both

sides (channel and shell) contain water as the working fluid. The water density, viscosity and the

conductivity were kept constant. The present simulation was carried out under steady state

condition. The flow is incompressible and the radiation is neglected.


Outlet: Heat Flux Outlet:
Zero Gage Zero Gage Heat Flux
Inlet: flow rate&
Temp.

symmetry
Inlet: flow rate&
Temp.

(a) Inner channel (b) Shell side (Outer) channel

Fig. 3. Schematic of CFD models.


(a) Mesh description

40
19 million, 0.005 mm
16 million, 0.01 mm
12.5 million, 0.01 mm
35 9 million, 0.03 mm
6 million, 0.08 mm
Temperature ( C)
o

30

25

20

15
0.0 0.2 0.4 0.6 0.8 1.0 1.2
X/L

(b) Mesh independence validation.

Fig. 4. Schematic of the CFD model and its validation. (with validation against temperature
variation in shell channel with β = 75°, Re = 912, and T inlet = 20 °C)

The schematic of the mesh description of shell channel is shown in Fig. 4(a), while Fig. 4(b)

shows the calculated fluid temperature in the middle distance between the upper and lower plate
at the symmetry surface subject to the number of meshes. The mesh-independence study was

conducted at a high Reynolds number of 912 to verify the turbulent flow conditions. As can be

seen in Fig. 4(b), no detectable variation in the measured temperature is seen when raising the

mesh element over 16 millions. Hence the mesh elements of 16 millions with a wall layer thickness

of 0.01 mm were used throughout the simulations for the shell channel. For the inner channel, the

mesh-independence condition prevails at 22 million elements with 0.01 mm wall layer thickness.

For further elaborations of the calculated results, the thermal and hydraulic performance of the

SAPHX are in terms of heat transfer performance (heat transfer coefficient and Nusselt number)

and frictional performance (pressure drop and friction factor) accordingly.

For the calculated channel pressure drop, ∆𝑃𝑐ℎ , the channel friction 𝑓𝑐ℎ factor is defined as:
∆𝑃𝑐ℎ × 𝐷ℎ (1)
𝑓𝑐ℎ =
2𝜌𝑉 2 𝐿𝑒𝑓𝑓

Where Dh is the hydraulic diameter,  density of working fluid, Leff the effective length in
plate. The mean velocity of channel 𝑉 is
𝑚̇𝑐ℎ (2)
𝑉=
𝜌𝐴𝑐ℎ

Where 𝑚̇𝑐ℎ is the mass flowrate in each channel, Ach is the cross-sectional area of the plate.

For conventional plate change (rectangular plate), the hydraulic diameter 𝐷ℎ is

𝐷ℎ = 2𝑏 (3)

Where b is the mean channel flow gap as shown in Fig. 2(b). For the present shell and plate

configuration, Dh is varying from place to place for its circular plate geometry. Hence the hydraulic

diameter must be evaluated based on the general definition of hydraulic diameter:


𝐴𝑐ℎ (4)
𝐷ℎ = 4
𝑃ℎ

Where Ph is the periphery. In the case of a rectangular plate, the width is constant alongside

the flow direction, while the width keeps decreasing when moving away from the center of the

circular plate and reaches zero at the edge of the plate. In this regard, the present study adopts an

equivalent width for considering the effect of plate width. The average width 𝑤𝑎𝑣𝑔 of the circular

plate is derived based on the assumption of the width containing an equal weight functionality at

any distance from the center:


𝑟
∫ 2√𝑟 2 − 𝑥 2 𝑑𝑥 𝜋𝑟 (5)
𝑤𝑎𝑣𝑔 = 0 =
𝑟 2

Where x is the displacement from the circler plate center ( 𝑥 = 𝑟 sin 𝜃) and r is the radius of

the plate. By substituting Eq. (5) into Eq. (4) yields

𝜋𝑟𝑏 (6)
𝐷ℎ = 𝜑𝜋𝑟
+𝑏
2

Where  is the enlargement factor representing the surface incensement of the corrugation

relative to flat plate. The Nusselt number (Nu) is given as:

ℎ𝐷ℎ (7)
𝑁𝑢 =
𝑘

Where h is the heat transfer coefficient. k is the thermal conductivity of working fluid. For

given plate surface temperature 𝑇𝑤 , fluid bulk temperature 𝑇𝑏 and heat transfer rate 𝑄, the heat

transfer coefficient h is

𝑄 (8)
ℎ=
𝐴(𝑇𝑤 − 𝑇𝑏 )
For further validation of the present simulation, Figs. 5(a) and 5(b) shows respectively the

comparison of Nu versus Re, and 𝑓𝑐ℎ versus Re between the simulation results and experimental

results by Muley and Manglik [23]. The simulation is conducted with identical PHE geometry with

their 30°/30° plate arrangement. As shown in the figure, the simulations are in line with

experimental measurements in both heat transfer and frictional characteristics. The maximum error

in Nu at low Reynolds number is about 20% and the difference is significantly reduced at higher

Reynolds number. In fact, in the case of high Reynolds number, the maximum deviation was less

than 11% while the mean error was about 7%. The difference falls in a rational range since the

viscosity is kept constant during the present simulation, and the influence of viscosity is lessened

when reducing the temperature difference between bulk and inlet (high Reynolds number).

Regarding 𝑓𝑐ℎ , the maximum deviation for the cases with low Reynolds number was 8% and the

deviation was about 6% in high Reynolds number.

70

60 Simulation results
Muley and Manglik [20]
50

40
Nu

30

20

10

0
0 500 1000 1500 2000 2500 3000 3500
Re

(a) Nu vs. Re
0.9

0.8 Simulation results


Muley and Manglik [26]
0.7

0.6

ch
f
0.5

0.4

0.3

0.2
0 500 1000 1500 2000 2500 3000 3500
Re

(b) 𝑓𝑐ℎ vs. Re

Fig. 5. Validation of the simulation results against experimental data for typical plate heat
exchanger. ( 𝛾 = 0.553, β = 30°/30°, 𝜑 = 1.29)

3. Results and Discussion

3.1 Temperature and velocity Profile

The geometric details of shell and plate channels used in this study is listed in Table 1. All the

models are considered symmetric, hence any two plates which construct one channel share one β.

Fig. 6 shows the infrared image of conventional plate heat exchanger [24] which is used as the

qualitative comparison in association with the present shell and plate geometry. Using the

operating conditions in Table 2, Fig. 7 (a) shows the temperature profile of the SAPHX at the

middle plane amid two plates for both inner and shell side channels with β = 45° and 75°

respectively, while Fig. 7(b) represents the velocity profile and Fig. 7(c) illustrates the velocity

vector at shell side channel.

As seen from the comparison between Fig. 6 and Fig. 7(a), for the conventional plate heat

exchanger, the peak temperature is more confined at the plate corner near the outlet port while for
the inner channel of SAPHX, the fluid containing the highest temperature is limited within the

brink region of exit port and the edge of the plate diameter. Yet the portion with higher temperature

for SAPHX is comparatively smaller than that in conventional plate heat exchanger. In addition,

it is found the plate with a larger chevron angle (β = 75°) contains a better temperature uniformity

when comparing with that of β = 45°. Yet the higher temperature region for β = 75° relative to β

= 45° at the exit is conspicuously reduced. This can be also made clear from the velocity profile

and velocity vector shown in Figs. 7(b) and 7(c) where the flowing water alongside the plate shows

more unidirectional for β = 75°, indicating a better flow distribution and a higher heat transfer

performance accordingly. From the temperature distribution of the present SAPHX, the variation

of temperature gradient is qualitatively less pronounced than that of PHX as depicted in Fig. 6(a).

Based on the studies by Marugán-Cruz et al. [25], Yapıcı and Albayrak [26], and Alzaharnah et

al. [27] regarding the thermal stress characteristics alongside a circler tube subject to a uniform

and non-uniform heat flux, higher thermal stress occurs as a consequence of larger temperature

gradient. Hence, the portion near the outlet/inlet port offers a strong ambience to develop high

thermal stress since the temperature at these part grows dramatically as clearly shown in Fig. 6 and

Fig. 7(a). Note that the effect of thermal stress concentration is comparatively relieved for the

present SAPHX configuration for its round plate geometry and the extra thickness to the plate

thickness near the port exit. On the other hand, at temperature variation in the shell channel of

SAPHX appears to be quite uniform (Fig. 7(b)), and is free of any conceivable temperature

fluctuation. Additionally, the free space containing at the outlet also prevents any sudden

accumulation of the hot fluid. To sum up, a lower thermal stress may prevail at the shell side for

its much gentler variations in temperatures, thereby reducing the leakage concerns and offering a

higher pressure operating conditions as far as long-term operational reliability is concern.


Table 1
Geometric details of inner and shell side channels.
Parameter β = 45ο β = 75ο
Plate thickness, t (mm) 0.8 0.8
Plate diameter, D (mm) 215 215
Port diameter, d (mm) 40 40
Mean channel flow gap, b (mm) 2.45 2.45
Corrugation pitch Pc (mm) 7.5 7.5
Mean channel flow area , Ach (m2) 0.000273 0.000273
Enlargement factor 𝜑 1.19 1.19
Port area, Aport (m2) 0.00126 0.00126
2
Plate circular area, Aplate (m ) 0.00126 0.00126
Single Plate heat transfer area, (m2) 0.0452 0.0452
Inner channel heat transfer area, 0.088 0.088
At,in(m2)
Shell channel heat transfer area, At,out 0.0904 0.0904
(m2)
Effective length, Leff,in (mm) 145 145
Effective length, Leff, out (mm) 215 215
Dh (mm) 2.5 2.5

Fig. 7(b) shows the velocity profiles. It can be noticed that the fluid-flow in β = 75ο channel

has a higher bulk velocity than that for the fluid with β = 45ο channel. The flowing fluid takes

straight passages while the flow of β = 45ο channel scatters in multiple directions. The difference

in flow pattern inside β = 45ο and β = 75ο channels can also be noticed in Fig. 7(c).

Fig. 6. Infrared image for the temperature distribution at the plate surface of the
conventional plate exchanger [24].
(c) velocity vector.
Fig. 7. Temperature and velocity profiles at the middle of the plate for both inner and
shell channel; (top) & shell- side (bottom) channels, for 45° (left) & 75°(right) chevron
angle.

Table 2
Operating conditions for the fluid flow.
Parameter Values

Channel mass flow rate 0.08 kg/s

Total heat rate from plates to the fluid 1600 W

Inlet fluid temperature 20 C

4.2 Friction factor and Nusselt number in both inner and shell channels

The variation of 𝑓𝑐ℎ versus Re for inner and shell channels for β = 45° and 75° is illustrated

in Fig. 8(a). 𝑓𝑐ℎ and Dh were obtained from Eq. (1) and Eq. (7) respectively, while the plate

diameter is regarded as the effective length. In Fig. 8(a), the calculated results are also compared

with those correlations calculated from Muley and Manglik [23], Kumar [28], and Thonon et al.

[29]. The calculations are generally in agreement with Muley and Manglik [23] and Kumar [28]

for the same β, while the reported data for shell side channel, β = 45° is more close to the correlation

proposed by Thonon whose chevron angle is 30° [29]. The calculated friction factor is generally
lower than the existing correlations especially when the Reynolds number is low. There are two

possible explanations regarding the differences. Firstly, the prior correlations were derived from

test results from conventional PHXs and there is geometric difference between these two heat

exchangers. Secondly, the hydraulic diameter for the present SAPHX is based on an equivalent

hydraulic diameter defined in Eq. (6) which may offer lower friction factors when comparing with

the conventional plate having a fixed Dh. Nevertheless, the tendency of the frictional behaviors for

the SAPHX and the conventional PHX looks quite alike. For a lower Reynolds number, the

frictional factor is decreased dramatically with the rise of Reynolds number and it becomes less

influenced when the Reynolds number is sufficiently high. In the meantime, for the present

SAPHX, it appears that the friction factors between inner and shell channel is rather small when 

= 75. However, the friction factors for inner channel having  = 45 are significantly higher than

those of  = 75 (around 60~95% higher). Notice that there is difference between the inner and

shell channel despite they share the same plate and therefore is expected to contain the same fluid

flow behavior. However, this is not the case for the present SAPHX when comparing with

conventional PHX. The flow in and out of the inner channel encounters a sharp turning (90) while

the flow in the shell channel is unidirectional without abrupt detouring, and the fluid flow in the

shell channel behaves like a cross-flow. The pressure drops in the inner side comprise both in/out

turning loss and frictional loss from the corrugated channels. For a high chevron angle like  =

75, the frictional pressure drop is in control, thereby the difference in friction for factors is small

for both chevron angles. Conversely, the in/out turning loss overpowers the frictional loss when

the  is reduced to 45, thereby resulting in an appreciably higher friction factor in the inner

channel.
Fig. 8(b) shows the variation of the Nusselt number with the Reynolds number, and the results

of Muley and Manglik [23], Kumar [28] and Roa et al. [30] are also illustrated for comparisons.

The results of Muley and Manglik [23] are in a good agreement with inner channel with β = 60.

The data of Rao et al. [30] is lower than the present study for the correlation is developed for β =

30°. However, the data of Kumar [28] for β = 45° offers a quit higher value than expected. This

discrepancy is associated with the dissimilarity in the corrugation parameters which can affect the

thermal performance, detailed elaboration will be addressed further in section 4.3.

For the heat transfer performance for the present SAPHX as seen from Fig. 8(b), the shell

channel with β = 75° contains the highest Nu and the corresponding enhancements becomes more

and more pronounced when the Reynolds number exceeds 150. Yet the Nu is about the same for

shell channels with both chevron angles when Re < 150. Hence, at a lower Re operating conditions,

β = 45° may be a better choice over β = 75° since the 𝑓𝑐ℎ of the shell and inner sides channels with

β = 75° are much higher than inner and shell side channels with β = 45°. For the comparison

between the shell channel and inner channel, at a low Re range, i.e. Re < 200, the Nu of shell

channel outperforms the inner channel with β = 45° as can be seen in Fig. 8(b). Conversely, Nu

for the shell-channel is lower than that of the inner channel in the range where Re > 200 and β =

45°, while 𝑓𝑐ℎ for shell channel is always lower than that of inner side. To explain this

phenomenon, one has to resort to the difference in flow pattern between inner and shell side as

addressed earlier. A schematic of the flow pattern in inner and shell channel is demonstrated in

Fig. 8(c). For the inner channel, the impingement of the fluid flow at the entrance port may spread

partial flowrate toward the exit port while also forcing some of the impinging flow onto the plate

heading backwards and forming re-circulation. Analogous flow pattern prevails at the exit port of

the inner channel. The recirculated flow pattern jeopardizes the heat transfer performance at the
inner channels. Note that this phenomenon becomes more prominent when the chevron angle is

increased. This is because the appreciable rise in friction contribution for the effective plate length

for heat exchange (see Fig. 8(c)) enforces more flow re-circulation at the inlet and exit ports,

thereby deteriorating the heat transfer performance further. On the other hand, the flow pattern in

the shell channel, as illustrated in the shell Fig. 8(c), shows no re-circulation with a unidirectional

behavior. In summary of the foregoing discussion, heat transfer performance for shell channel is

further reinforced with a high chevron angle of 75 and high Reynolds number when comparing

to inner channels.

2.0
45o inner channel
1.8 45o shell side channel
75o inner channel
1.6
75o shell side channel
1.4 Muley & Manglik, 60o [20]
o
Kumar, 45 [25]
1.2 Thonon et al., 30o [26]
fch

1.0

0.8

0.6

0.4

0.2

0.0
0 200 400 600 800 1000

Re
(a) 𝑓𝑐ℎ vs. Re.
70

60

50

40
Nu

30 45o inner channel


45o shell side channel
20 75o inner channel
75o shell side channel
Rao et al. 30o [28]
10
Muley & Manglik, 60o [20]
o
Kumar, 45 [25]
0
0 200 400 600 800 1000
Re

(b) Nu vs. Re.

(c) Schematic of flow patterns in shell and inner channels.

Fig. 8. Hydraulic and thermal performance of inner and shell side channels for
different chevron angles.

4.3 Effect of corrugation parameters


To investigate the effect of the corrugation pitch, Pc and mean channel gap, b, on the thermal

performance inside the channel, the corrugation aspect ratio, , which is termed as 2𝑏⁄𝑃𝑐 [22] is

introduced and examined in association with the Reynolds number for both inner and shell channel

with β = 45°. Doo et al. [31] studied the same effect using 𝑃𝑐 ⁄𝑏 ratio while the study by Zhicheng

et al. [12] was based on . Fig. 9 (a) illustrate Nu verses  at three different Reynolds number (47,

364, and 912, respectively) for the inner channel with β = 45°. It can be seen that Nu is increased

moderately with  up to 0.75, and starts to decline for a further rise of . Similar trends prevail for

these three Reynolds number. Note that the maximum variation in Nu in the range of 0.38 <  < 1

is 6% for Re = 47, and is 11% for Re = 912. The tendency regarding the influence of  is in a good

agreement with [12, 31] who reported 10-12% improvement in heat transfer performance for

typical channels within PHE having similar  range to the present study. Correspondingly, Fig.

9(b) illustrates Nu verses  at identical three Reynolds number for the shell channel with β = 45°.

The optimum  value for the highest Nu is 0.53, meaning that shell channel performs better at a

smaller b when comparing with the inner channel. The variation trend of Nu for different Reynolds

number is also similar. However, in the range of 0.38 <  < 1, the maximum variation in Nu for is

just 4% for Re = 47, while it is 9% for Re = 912. Hence, it is concluded that the heat transfer

performance in the shell channel is less influenced by  when comparing with the inner channel.

This is somehow expected for more uniform flow pattern in the shell channel is encountered.

The local heat transfer coefficient profile at the plate surface inside the inner channel and shell

side channel is respectively illustrated in Figs. 10 (a) and 10(b) to explain the influence of  on

heat transfer performance and the reasons of occurrence of optimum . It can be noticed from Fig.

10(a) and 10(b) that the local peak heat transfer coefficient occurs at the trough’s and crest’s of the

plate corrugation of the inner and shell channel, respectively. As b decreases, the associated surface
area of crest and trough which containing the peak heat transfer coefficient, also increases

accordingly. As a consequence, the average heat transfer coefficient (h and Nu) is increased.

However, the hydraulic diameter Dh, reduces as b is decreased, therefore imposing an adverse

affect on Nu. At first sight, it looks contradictory to the previous statement. This contradiction can

be explained with help of optimum . Until the optimum  value is reached, the influence of Dh on

Nu dominates over h. However, once beyond the optimum , the surface area of the crest and

trough region reduces dramatically and impairs the heat transfer coefficient significantly. In

summary of these two competitive effects, the effect of deteriorating heat transfer coefficient

surpasses the contribution of Dh after the optimum .

60 50

45
50
Re 47
Re 364 40 Re 47
40 Re 912 Re 364
35
Nu / Nureff

Re 912
Nu

30 30

20 25

20
10
15
0
10
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
(2b / Pc)
(2b / Pc)

(a) Inner channel flow (a) Shell side channel flow

Fig. 9. Effect of  on normalized Nu for inner and shell side channels with β 45° subject to
Reynolds number of 47, 364, and 912.
Wm-2 K-1
trough crest
crest

trough

(a) inner channel (b) shell side channel

Fig. 10. Local heat transfer coefficient distribution alongside the plate surface for inner and
shell channels. (V = 0.22 m s-1, Q = 1200 W, Tinlet = 20 °C)
0.4

45o
0.3 75o
PHX )/ PHX

0.2
SAPHX -

0.1

0.0

0 200 400 600 800 1000

Re
(a) Improvement in energy conversion efficiency vs. Re of SAPHX relative to PHE.
0.7

(P pumping _PHX - P Pumping_ SAPHX )/ P pumping _PHX


0.6

0.5
45o
75o
0.4

0.3

0.2

0.1
0 200 400 600 800 1000
Re
(b) Energy saving in required pumping power vs. Re of SAPHX relative to PHE.
Fig. 11. Capability of thermal energy conversion and the energy saving in required pumping vs.
Reynolds number (shell side) between the proposed shell and plate heat exchanger and plate heat
exchanger. (Effective temperature difference = 10 °C, water in the channel side and ammonia at
the shell side, both plates contain the same heat transfer area).

For further elaboration of the superior heat transfer performance of the shell and plate heat

exchanger, the optimum design of the present simulation is used to calculated the thermal energy

conversion subject to a low-temperature OTEC operating condition where water is used in the

channel side while ammonia is given in the shell side with the effective mean temperature

difference of 10 C. The comparison is made with the counterpart heat exchanger – PHE having

the same surface area. Calculation is made with chevron angles of 45 and 75, respectively. Note

that the comparison is based on shell side. And the performance is in terms of ratio of thermal

energy conversion effectiveness () between SAPHX and PHE as shown in Fig. 11(a). As depicted

in Fig. 11(a), the thermal energy conversion effectiveness for SAPHX is always higher than its

PHE counterpart irrespective of the operating conditions and the chevron angles. Yet, the
difference is especially prominent when the chevron angle is raised to 75. The thermal energy

conversion effectiveness for SAPHX exceeds those of PHE by 12-35% throughout the test range.

Yet, the energy saving associated with the pumping power for the SAPHX relative to PHE, is

given in Fig. 11(b). Note that the pumping power, P, is calculated from the product of pressure

drop across the plate and the associated volumetric flowrate. The energy saving in the pumping

power for the optimum SAPHX design when compared with PHX is even more pronounced. For

a higher chevron angle of 75, the energy saving of the required pumping power ranges from 20%

to 30%. Yet the energy saving in pumping power can be raised substantially to 64% at a lower

chevron angle of 45. In summary of the forgoing discussion, the proposed shell and plate heat

exchanger not only contains superior thermofluids characteristics but also offers significant

thermal energy conversion effectiveness and an appreciable energy saving in required pumping

power. The proposed SAPHX design shows far more prominent feature than the PHE as far as

low-temperature thermal energy conversion is concerned.

5 Conclusions

High-efficiency thermal energy conversion is pivotal for low-temperature difference

applications such as Ocean Thermal Energy Conversion (OTEC). The shell and plate heat

exchanger evolved from its predecessors (shell-and-tube heat exchanger and plate heat exchanger)

is regarded as the best candidate for featuring both superior performance and robust configuration.

The present study examined the hydraulic and thermal performance of the shell and plate heat

exchangers through detailed numerical investigations. The working fluid used for the study is

water under steady state and single-phase conditions. The corresponding chevron angles are β =

45 and 75, respectively. And the Reynolds number ranges from 47 to about 1000. Based on the

foregoing discussions, the following conclusions are drawn:


1. The simulations are validated and in line with the experimental data of conventional plate

heat exchangers. Unlike those of conventional plate heat exchanger, the heat transfer and frictional

performance differ significantly between shell and inner channel.

2. The friction factors for the shell and plate channels are normally lower than the

conventional rectangular plate channel for its varying hydraulic diameter.

3. For a larger chevron angle (75), the friction factor in the shell and inner channel is

comparable. However, the friction factor in the shell channel is much lower than that in inner

channel at a smaller chevron angle (45). The major difference is due to the dramatic difference of

the flow pattern in/out the inner and shell channel. The shell channel contains more unidirectional

flow pattern while some flow re-circulation may occur at the entrance/exit port of the inner channel.

4. The flow pattern and temperature distribution in the shell channel are generally more

uniform than those in inner channel.

5. The shell channel with β = 75° contains the highest Nu and the corresponding

enhancements becomes more and more pronounced when the Reynolds number exceeds 150. At

a lower Reynolds number operating condition, a better overall performance is seen for a lower β

= 45° due to its much lower friction factor.

6. The heat transfer performance for shell channel exceeds the inner channel appreciably

when the chevron angle is high (75) and operated at a comparatively high Reynolds number.

7. The heat transfer performance depends on the corrugation aspect ratio, and an optimum

ratio is encountered for the present shell and plate heat exchanger. For the inner channel, the

optimum value is around 0.75 while it is 0.53 for the shell channel.

8. The local peak of the heat transfer coefficient occurs at the trough’s and crest’s of the plate

corrugation of the inner and shell channel, respectively.


9. The optimum design of the present simulation is used to simulate the thermal energy

conversion effectiveness subject to a low-temperature OTEC operating condition with the effective

mean temperature difference of 10 C. It is found that the thermal energy conversion effectiveness

for SAPHX exceeds those of PHE by 12-35%. Yet the energy saving in the required pumping

power is even more prominent, ranging from 20%-64%.

Acknowledgements

The authors acknowledge the financial support from ministry of science and technology,

Taiwan under contract Nos. 108-2622-E-009 -027-CC2 and 108-2811-E-009 -539.

Nomenclature
A: single plate heat transfer area, [m2]
Ach: mean channel flow area, [m2]
Aplate: circle plate area, [m2]
Aport: port cross section area, [m2]
Avg.: Average value
b: Mean channel flow gap, [mm]
D: plate diameter, [mm]
Dh: Hydraulic diameter, [mm]
d: port diameter, [mm]
𝑓𝑐ℎ : channel friction factor
h: heat transfer coefficient, [Wm-2 K-1]
HX: heat exchanger
k: fluid conductivity, [W m-1 K-1]
Leff: effective length, [mm]
𝑚̇: mass flow rate, [kg s-1]
𝑁𝑢: Nusselt number
P: required pumping power to flow across the shell side [W]
Ph: perimeter [m]
∆𝑃𝑐ℎ : channel pressure drop, [Pa]
PHE: plate heat exchanger
Q: heat transfer rate, [W]
r: plate radius, [mm]
𝑅𝑒: Reynolds number
SAP: shell and plate heat exchanger
Tb: fluid bulk temperature, [°C]
𝑇𝑤 : near wall temperature, [°C]
t: plate thickness, [mm]
U: overall heat transfer coefficient, [Wm-2 K-1]
𝑉: mean channel flow velocity, [m s-1]
w: width, [mm]
Greek letters
𝛽: chevron angle, [°]
: thermal energy conversion effectiveness
𝜑: enlargement factor
μ: fluid dynamic viscosity, [Pa s]
𝜌: density, [kg m-3]
𝛾: corrugation profile aspect ratio 2𝑏⁄𝑃𝑐

References
[1] Z. Ma, Y. Wang, S. Wang, Y. Yang, Ocean thermal energy harvesting with phase change material for
underwater glider, Applied Energy, 178 (2016) 557-566.
[2] T. Nakaoka, H. Uehara, Performance test of a shell-and-plate-type condenser for OTEC, Experimental
Thermal and Fluid Science, 1 (1988) 275-281.
[3] N. Tran, C.-C. Wang, Effects of tube shapes on the performance of recuperative and regenerative heat
exchangers, Energy, 169 (2019) 1-17.
[4] S. Kakaç, H. Liu, A. Pramuanjaroenkij, Heat exchangers: selection, rating, and thermal design, CRC
press, 2002.
[5] M.M. Abu-Khader, Plate heat exchangers: Recent advances, Renewable and sustainable energy
reviews, 16 (2012) 1883-1891.
[6] T.M.A. Elmaaty, A. Kabeel, M.J.R. Mahgoub, S.E. Reviews, Corrugated plate heat exchanger review, 70
(2017) 852-860.
[7] J. Zhang, X. Zhu, M.E. Mondejar, F. Haglind, A review of heat transfer enhancement techniques in plate
heat exchangers, Renewable and Sustainable Energy Reviews, 101 (2019) 305-328.
[8] C. Gulenoglu, F. Akturk, S. Aradag, N.S. Uzol, S. Kakac, Experimental comparison of performances of
three different plates for gasketed plate heat exchangers, International Journal of Thermal Sciences, 75
(2014) 249-256.
[9] J. Yang, A. Jacobi, W.J.A.T.E. Liu, Heat transfer correlations for single-phase flow in plate heat
exchangers based on experimental data, 113 (2017) 1547-1557.
[10] W. Yoon, J.H.J.I.J.o.H. Jeong, M. Transfer, Development of a numerical analysis model using a flow
network for a plate heat exchanger with consideration of the flow distribution, 112 (2017) 1-17.
[11] C. Turk, S. Aradag, S.J.I.J.o.T.S. Kakac, Experimental analysis of a mixed-plate gasketed plate heat
exchanger and artificial neural net estimations of the performance as an alternative to classical
correlations, 109 (2016) 263-269.
[12] Y. Zhicheng, W. Lijun, Y. Zhaokuo, L. Haowen, Shape optimization of welded plate heat exchangers
based on grey correlation theory, Applied Thermal Engineering, 123 (2017) 761-769.
[13] K. Song, T. Tagawa, Z. Chen, Q. Zhang, Heat transfer characteristics of concave and convex curved
vortex generators in the channel of plate heat exchanger under laminar flow, International Journal of
Thermal Sciences, 137 (2019) 215-228.
[14] P. Jamzad, J. Kenna, M. Bahrami, Development of novel plate heat exchanger using natural graphite
sheet, International Journal of Heat and Mass Transfer, 131 (2019) 1205-1210.
[15] M. Elias, R. Saidur, R. Ben-Mansour, A. Hepbasli, N. Rahim, K. Jesbains, Heat transfer and pressure
drop characteristics of a plate heat exchanger using water based Al 2 O 3 nanofluid for 30° and 60° chevron
angles, Heat and Mass Transfer, 54 (2018) 2907-2916.
[16] M.B. Kim, C.Y. Park, An experimental study on single phase convection heat transfer and pressure
drop in two brazed plate heat exchangers with different chevron shapes and hydraulic diameters, Journal
of Mechanical Science and Technology, 31 (2017) 2559-2571.
[17] T.S. Khan, M.S. Khan, Z.H. Ayub, Single-phase flow pressure drop analysis in a plate heat exchanger,
Heat transfer engineering, 38 (2017) 256-264.
[18] E. Özkaya, S. Aradag, S. Kakac, CFD Aided Design of Heat Transfer Plates for Gasketed Plate Heat
Exchangers, in: ASME 2014 12th Biennial Conference on Engineering Systems Design and Analysis,
American Society of Mechanical Engineers, 2014, pp. V003T012A004-V003T012A004.
[19] L.O. Freire, D.A.d. Andrade, On applicability of plate and shell heat exchangers for steam generation
in naval PWR, Nuclear Engineering and Design, 280 (2014) 619-627.
[20] O.P. Arsenyeva, L.L. Tovazhnyanskyy, P.O. Kapustenko, G.L. Khavin, A.P. Yuzbashyan, P.Y. Arsenyev,
Two types of welded plate heat exchangers for efficient heat recovery in industry, Applied Thermal
Engineering, 105 (2016) 763-773.
[21] J. Lim, K.S. Song, D. Kim, D. Lee, Y. Kim, Condensation heat transfer characteristics of R245fa in a shell
and plate heat exchanger for high-temperature heat pumps, International Journal of Heat and Mass
Transfer, 127 (2018) 730-739.
[22] A. Muley, R. Manglik, Experimental investigation of heat transfer enhancement in a PHE with β= 60
chevron plates, in: Proceedings of the Second ISHMT-ASME Heat and Mass Transfer Conference and
Thirteenth National Heat and Mass Transfer Conference, Dec, 1995, pp. 28-30.
[23] A. Muley, R. Manglik, Experimental study of turbulent flow heat transfer and pressure drop in a plate
heat exchanger with chevron plates, Journal of heat transfer, 121 (1999) 110-117.
[24] S. Jin, P. Hrnjak, Effect of end plates on heat transfer of plate heat exchanger, International Journal
of Heat and Mass Transfer, 108 (2017) 740-748.
[25] C. Marugán-Cruz, O. Flores, D. Santana, M. García-Villalba, Heat transfer and thermal stresses in a
circular tube with a non-uniform heat flux, International Journal of Heat and Mass Transfer, 96 (2016)
256-266.
[26] H. Yapıcı, B. Albayrak, Numerical solutions of conjugate heat transfer and thermal stresses in a circular
pipe externally heated with non-uniform heat flux, Energy Conversion and Management, 45 (2004) 927-
937.
[27] I. Alzaharnah, M. Hashmi, B. Yilbas, Thermal stresses in thick-walled pipes subjected to fully
developed laminar flow, Journal of materials processing technology, 118 (2001) 50-57.
[28] H. Kumar, The plate heat exchanger: construction and design, in: Institute of Chemical Engineering
Symposium Series, Vol. 86, 1984, pp. 1275-1288.
[29] B. Thonon, R. Vidil, C. Marvillet, Recent research and developments in plate heat exchangers, Journal
of Enhanced Heat Transfer, 2 (1995).
[30] B.P. Rao, B. Sunden, S.K. Das, An experimental and theoretical investigation of the effect of flow
maldistribution on the thermal performance of plate heat exchangers, Journal of heat transfer, 127 (2005)
332-343.
[31] J. Doo, M. Ha, J. Min, R. Stieger, A. Rolt, C. Son, Theoretical prediction of longitudinal heat conduction
effect in cross-corrugated heat exchanger, International Journal of Heat and Mass Transfer, 55 (2012)
4129-4138.
Highlights
– Thermofluids performance of shell and plate heat exchanger (SAPHX) is reported
numerically.
- SAPHX contains quite different flow pattern than the conventional plate heat
exchanger.
- Performance in shell side is normally superior to inner channel side.
- The difference in shell side and inner channel is explained thoroughly.
- The optimum ratio between the plate corrugations is investigated in details.
Submitted to Applied Thermal Engineering
Authors: Ali Abbas, Howard Lee, Akash Sengupta and Chi-Chuan Wang*
_____________________________________________________________________

I am the corresponding author (Chi-Chuan Wang) and declare no conflict of interest of this
submission.

You might also like