You are on page 1of 14

Digital Chemical Engineering 3 (2022) 100014

Contents lists available at ScienceDirect

Digital Chemical Engineering


journal homepage: www.elsevier.com/locate/dche

Original Article

Computational Fluid Dynamics (CFD) analysis of the heat transfer and fluid
flow of copper (II) oxide-water nanofluid in a shell and tube heat exchanger
Patricia Anne D. Cruz a, Ed-Jefferson E. Yamat a, Jesus Patrick E. Nuqui b,∗, Allan N. Soriano b,∗
a
School of Chemical, Biological, and Materials Engineering and Sciences, Mapúa University, Intramuros, Manila, Philippines
b
Chemical Engineering Department, Gokongwei College of Engineering, De La Salle University, 2401 Taft Avenue, Manila, Philippines

a r t i c l e i n f o a b s t r a c t

Keywords: Numerical analysis of the thermal and flow behavior of CuO-water nanofluid under turbulent regions in a shell
CuO-water nanofluid and tube heat exchanger was conducted using ANSYS Fluent. Twenty-nine (29) nm diameter CuO nanoparticles,
Heat transfer coefficient and water as base fluid were used in the study. The nanofluid was simulated at different particle loading (0.1 to
Performance index
1%vol), and under three sets of Reynolds number (ranging from 17,000 to 71,000), to study the effects on heat
Pressure drop
transfer coefficient, pressure drop, and nanofluid thermal and hydrodynamic behavior. Increasing the particle
Reynolds number
Shell and tube heat exchanger loading and Reynolds number was found to enhance both the heat transfer rate and pressure drop. A maximum
of 48% enhancement in the heat transfer was observed at the highest particle loading, but with the consequence
of doubled pressure drop. Performance indices greater than 1 were attained for particle loading below 0.25%vol,
regardless of the Reynolds number. The conditions that produced the highest index were at the lowest particle
loading and lowest Reynolds number. No significant difference in the flow behavior between water and CuO-
water nanofluid was observed. However, the thermal profiles for 0.1%vol CuO-water nanofluid highlighted the
enhancements in heat transfer along the shell and tube heat exchanger.

1. Introduction are conventionally used as coolants, such as oil, water and ethylene gly-
col. These additives, which are usually less than 5 volume (vol %) per-
A heat exchanger is a device that facilitates the transfer of energy cent, were found to boost the thermophysical properties of the working
between one or more fluid streams with an existing temperature dif- fluid (Haghighi, 2015), thus are solid candidates for meeting high heat
ference. Chemical process industries such as power generation, petro- transfer requirements.
chemical, and heat recovery systems utilize heat exchangers as a means The conventional techniques for heat transfer enhancement are made
of providing effective and suitable heat transfer for cooling, heating, by either altering the interacting parameters or testing a better perform-
and material phasing change, depending on the industry’s nature (Devi ing coolant. Studies have concluded that the former method, design
& Nagamani, 2015). Shell and tube heat exchangers are indirect contact modification has almost been maximized, thus increasing the interest
devices wherein heat transfer is achieved as the fluids pass separately on utilizing the benefits of nanofluids as coolants. Initial research am-
through the shell and tube bundles, by means of conduction and con- plified the growing potential of nanofluids, with studies focusing on the
vection (Othman, 2009). These heat exchangers are designed to achieve thermophysical properties, rheology, and flow characterization. Most
optimal heat load through a range of flexible variations. Geometric pa- of these emphasized that among the drawbacks of utilizing NFs in heat
rameters (e.g. diameter, length, number of tubes, etc.) may be modified transfer are the effects on pressure drop due to changes in viscosity.
to increase the surface area for heat transfer, while fluid flow signif- Although the thermal conductivity increase is proven in many studies,
icantly affects performance parameters such as the heat transfer rate minimal changes in viscosity are difficult to achieve. Such condition is
and effectiveness (Jadhav & Koli, 2014). Over the years, the demand significant since a large increase in viscosity is found to inhibit the heat
for high heat transfer rates at minimal pressure drop remained, while transfer within the equipment (Pantzali et al, 2009). Studies on water
technological developments are aimed towards reduction of energy con- systems of titanium (IV) dioxide (TiO2 ), copper (II) oxide (CuO), and
sumption and costs for these units. Innovations in heat transfer enhance- carbon nanotubes, among others, emphasized that the flow regime at
ment techniques are now advancing towards the use of fluid additives which the fluid flows is as important as its thermophysical properties.
such as nanofluids (NFs), which are colloidal distribution of nanoparti- Conditions for achieving only a marginal viscous effect on the system
cles (e.g. metals, oxides, CNT or carbon nanotubes) in base fluids that are given high importance in order to monitor and control the pres-


Corresponding authors.
E-mail addresses: jesus_nuqui@dlsu.edu.ph (J.P.E. Nuqui), allan.soriano@dlsu.edu.ph (A.N. Soriano).

https://doi.org/10.1016/j.dche.2022.100014
Received 28 September 2021; Received in revised form 14 February 2022; Accepted 15 February 2022
2772-5081/© 2022 The Authors. Published by Elsevier Ltd on behalf of Institution of Chemical Engineers (IChemE). This is an open access article under the CC
BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/)
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

sure drop (Pantzali et al, 2009). As such, the ratio between the heat
transfer improvement and the pressure drop rise is then adapted as
the basis for evaluating nanofluid effectiveness. This notion also led to
the experimental determination of the optimum concentration levels for
nanofluid loading in heat exchangers. The use of nanofluids in enhanc-
ing the design of heat transfer systems is equally essential and critical in
nature, typically requiring time and economic considerations (Skocilas
& Palaziuk, 2015). Nanofluid preparation, property evaluation, and the
actual testing have been proven to be lengthy and overwhelmingly ex-
pensive. In addition, the sole implementation of experimental and ana-
lytical methods such as the Bell-Delaware, Logarithmic Mean Temper-
ature Difference (LMTD), and Number of Transfer Units (NTU) method
are iteratively complex. These methods may be reduced through reso-
lution of the physical system to computer models and simulations, such
as application of a Computational Fluid Dynamics (CFD) software like
ANSYS (Bhutta et al, 2011). CFD simulations provide engineers with
virtual modeling tools and quality solutions for evaluation of heat ex-
changer systems without the actual time, expenses and conditions re-
quired (Skocilas & Palaziuk, 2015). Thus, simulation provides a ground
for cost-effective testing and development of new equipment and pro-
cesses, such as the emerging application of nanofluids in heat transfer
systems.
Since the field of nanotechnology is relatively new, no concrete and Fig. 1. Task distribution in a CFD process (Masilungan-Manuel et al, 2015).
unified conclusion can be drawn from the open literature regarding the
extent of the overall enhancement caused by nanofluids on heat trans-
fer systems (Jokar & O’Halloran, 2013). Studies that utilized such flu- intensive industries with minimal offsets. Numerical results for the char-
ids as coolants for plate and tubular heat exchangers found a range of acterization of nanofluid behavior would help in determining the suit-
improvement on heat transfer. It is significant to note, however, that able operating conditions for its flow on this type of heat exchanger.
majority of the works focused on the determination of parameters such Additionally, the study also serves as a contribution to the research com-
as fluid outlet temperature, pressure drop, and flow regime conditions, munity during the onset of studies utilizing computer simulations.
and that analyses of the transport phenomena within the system was The sole nanoparticle selected for this study was copper oxide (CuO),
not available. Due to the instability of nanofluids and the high cost of which was infused with water as the base fluid. The nanofluid was sim-
experimentation and testing, the thermal conductivity and heat transfer ulated at different particle loading (0.1 to 1 vol %), under three sets of
coefficient data found in literature are inconsistent (Puliti et al., 2011). Reynolds numbers which correspond to the mass flow rates of water (0.5
However, this challenge sparked the interest of many researchers, hence to 1 kg/s), through ANSYS Fluent. The shell and tube heat exchanger
the growth of ongoing experimental studies for nanofluid property de- was modeled under steady-state and well-isolated (no heat losses) con-
termination and heat transfer enhancement. Further, since most of the ditions. The research undertakings involved single-phase approach in
studies were conducted using laboratory or fabricated set-ups, profil- modeling the nanofluid, such that the nanoparticles were assumed to
ing of the system, which included the flow and thermal behavior of the be perfectly homogenized in the base fluid. The flow regime for the
fluid were seldom presented. In the same manner, CFD simulations that model was assumed turbulent, based on the model used for validation.
specifically utilized nanofluids in shell and tube heat exchangers were Heat transfer is assumed to be achieved only through conduction and
fewer versus applications in other heat exchange equipment. Some theo- convection; all forms of radiation were considered negligible. Particle
retical studies on the optimum particle loading for plate heat exchangers mechanisms and interactions such as aggregation, sedimentation, foul-
recognized a few adaptations (Tiwari et al, 2015), while application of ing effect, and other disadvantages accompanied by the nanofluid were
such models to process industries had not yet been fully explored. Thus, not part of the study. The analysis was limited only to the shell side,
with these in mind, the present study developed a means of profiling the where the nanofluid will flow. The results of the present work pertain-
fluid dynamic behaviour of a nanofluid (based fluid containing nanopar- ing to the performance were compared to available literature whenever
ticles), which to our knowledge is still lacking. possible but for those other parameters (such as film coefficient) that
To this end, the study aims to develop a computational analysis of cannot be assessed due to lack of available data were not compared.
the heat transfer and fluid flow in a pilot-scale shell and tube heat ex-
changer using ANSYS 15.0 Fluent software. Specifically, it determines
the extent of the heat transfer enhancement using copper (II) oxide- 2. Methodology
water (CuO-water) nanofluid as the cold fluid. The performance of CuO-
water nanofluid was assessed by conducting a parameter study to mon- The computational fluid dynamic analysis of any system is comprised
itor the transport variables, performance index, and the thermal and of three general simulation tasks: pre-processing, processing, and post-
hydrodynamic flow profiles, which served as the criteria for determin- processing (Andersson, 2012; Masilungan-Manuel et al., 2015). Fig. 1
ing the effects of operating conditions variation, such as nanofluid flow represents a schematic diagram of the CFD process applied in this pre-
rate and particle loading. A comparative analysis of the heat transfer vent work. It begins with the pre-processing stage where the shell and
properties of the nanofluid and the base fluid were also conducted. tube heat exchanger domain is modeled by defining and generating its
Heat exchanger optimization techniques remain effective yet con- geometry and corresponding mesh. It is then to be followed by the pro-
ventional, and while these methods are considered practical for indus- cessing stage where the governing equations, numerical models, mate-
trial applications, the need for an excellent thermo-economic design rials, and the boundary conditions for the problem are defined. The so-
that will meet the fast-growing demand for heat exchange systems is lution calculations are then initiated based from the defined set-up. The
inevitable. The significance of this study lies in the exploration of the calculated results such as fluxes and thermal profiles can then be auto-
potential of nanoparticles as fluid additives, which may possibly be a matically viewed and gathered at the post-processing stage. The three-
key for efficiently meeting the high demand of heat transfer in energy- step process is repeated with varying boundary conditions and process

2
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Fig. 2. Pilot-scale shell and tube heat exchanger geometry.

Table 1
STHE geometry.

Variables Dimension

Heat Exchanger length 600 mm


Outer diameter (Tube) 20 mm
Outer diameter (Shell) 90 mm
No. of tubes 7
Number of baffles 6
Central Baffle spacing 86 mm
Tube bundle geometry Triangular
Pitch 30 mm
Baffle inclination angle 0o
Nozzle Diameter 36 mm Fig. 3. Meshing of the STHE–(A) along the length; (B) cross-sectional view; and
(C) tube mesh details.

parameters to come up with a set of results that can be compared and


the coolant Table 2. shows the results of validation of the current CFD
analyzed.
model using water as coolant. As presented in this table, the validation
The programs that were used for the three-step process for the study
results is satisfactory as shown by an overall percentage deviation of
were the built-in software available in ANSYS 15.0 except for the ge-
3.90 and 2.74 %, for heat transfer rate and temperature, respectively.
ometry creation. The initial algorithm of the CFD analysis was adapted
The resulting mesh of the STHE is given in Fig. 3 while the generated
from the computational analysis of a small shell and tube heat exchanger
mesh quality statistics are as follows–(a) node is 407,853 and (b) ele-
by Ozden and Tari (2010). ANSYS-Fluent was used as the solver for the
ments is 1,194,855.
determination of the heat transfer enhancement and effect on pressure
drop of CuO-water nanofluid as coolant.
2.2. Processing
2.1. Pre-Processing
Governing equations. Most of the equations, parameters and models
to be discussed in this section are adapted from different CFD studies of
Geometry. The shell and tube heat exchanger (STHE) geometry
shell and tube heat exchangers. For the governing equations, the time-
was adapted from the experimentally validated work of Ozden and
dependent terms of the conservation of mass, momentum (x, y, and z
Tari (2010), which is designed based on the standards of the Tubular
planes) and energy are dropped for the steady-state analysis of the STHE,
Exchanger Manufacturers Association (TEMA). The simplified heat ex-
as shown in Eqs. (1) to (6). The conservation of energy equation accounts
changer, as shown in Fig. 2, is a pilot-scale model that excludes the
for a dissipation function (Φ), as shown below (Ozden, 2010). Note that
shell covering and its flange. The three-dimensional model was built us-
all of the equations presented in this section are built-in functions of the
ing Rhino 3D and was later imported into the ANSYS Workbench design
ANSYS CFX solver for running the simulations.
module for further processing. The overall geometry and design param-
eters of the 1-1 STHE are summarized in Table 1. Also, for a simpli- ⃖⃖⃖⃗ = 0
∇(ρV) (1)
fied analysis of the shell-side, the tubes are modeled only as solid cylin-
ders. The simplified modeling approach allows analysis of the shell-side, 𝜕p 𝜕 τxx 𝜕 τyx 𝜕 τzx
where the nanofluid will flow. This modeling approach is adapted from ⃖⃖⃖⃖⃖⃗) = −
∇(ρuV + + + (2)
𝜕x 𝜕x 𝜕y 𝜕z
Ozden and Tari (2010) since the later results will be used to validate the
preliminary simulation using the based fluid which is water. 𝜕p 𝜕 τxy 𝜕 τyy 𝜕 τzy
⃖⃖⃖⃖⃖⃗) = −
∇(ρvV + + + + ρg (3)
Meshing. Computational meshing was done using ANSYS’s default 𝜕y 𝜕x 𝜕y 𝜕z
meshing client, ICEM CFD. The elements of the mesh for the entire
model are a mixture of tetrahedral and hexahedral cells; these will 𝜕p 𝜕 τxz 𝜕 τyz 𝜕 τzz
⃖⃖⃖⃖⃖⃖⃗) = −
∇(ρwV + + + (4)
serve as the data points for the temperature and flow of the fluids dur- 𝜕z 𝜕x 𝜕y 𝜕z
ing the numerical simulation. The mesh elements and nodes were scat- ( )
tered throughout the equipment using various combination of automatic ⃖⃖⃖⃖⃖⃗) = −p∇ V
∇(ρeV ⃗ + ∇(k∇T) + q + Φ (5)
meshing methods found in the program. Face and body selection was
used to control and concentrate the mesh near the wall regions. Rele-
[ [ )2 ( ( ] ( )2
vance of 25 and coarse curvature sizing as the relevance center were ( )2 )
𝜕u 𝜕v 𝜕w 2 𝜕u 𝜕v
used. To ensure that the applied meshing (and other conditions) are Φ= μ 2 ++ + +
𝜕x 𝜕y 𝜕z 𝜕y 𝜕x
acceptable, the results using the considered conditions (based on rele-
( )2 ( )2 ] ( )2
vance value and curvature sizing) were compared to the results obtained 𝜕u 𝜕w 𝜕v 𝜕w ⃗
+ + + + + λ ∇V (6)
by the work of Ozden and Tari (2010) in which pure water is used as 𝜕z 𝜕x 𝜕z 𝜕y

3
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Table 2
Validation of the current CFD model using water as coolant.

CFD analysis results

Ozden and Tari (2010) Current model Percentage difference

Mass flow rate(kg/s) Heat transfer rate (W) Tshell-out (K) Heat transfer rate (W) Tshell-out (K) Heat transfer rate (W) Tshell-out (K)

0.5 71,808 334.20 72,077.56 334.48 0.37 0.46


1 118,515 327.72 124,484.19 329.74 4.91 3.62
2 219,733 325.74 234,263.34 327.97 6.40 4.14

Table 3 Table 4
Realizable k-𝜀 model constants (ANSYS, 2013). Effective base fluid (Water)a and CuO-water nanofluidb properties used at
300 K.
Constants Values
Particle loading(vol %) 𝜌 (kg/m3 ) Cp (J-kg-K) 𝜇 (kg/m-s)
C1ε 1.44
C2 1.90 0.00 997.01 4179.00 0.000855
σ𝐤 1.00 0.10 1003.70 4158.47 0.001043
σε̃ 1.20 0.25 1011.96 4123.66 0.001104
0.50 1025.71 4066.89 0.001213
1 1053.22 3956.53 0.001449

Numerical modelling & turbulence models. Fluid flow within the a


Properties taken from Incropera (1996).
b
STHE is highly turbulent based on the preliminary simulations con- Properties calculated using Eqs. (8) to (10).
ducted by Ozden and Tari (2010). The choice of turbulence models is
essential to take account the different degrees of velocity, length and
time associated with flow instability within the STHE. These are approx- mal conductivity was defined using a piecewise-linear function. The said
imated based on the Navier-Stokes equations, which include the Large properties of the base fluid (water) are obtained from the data base of
Eddy Simulation (LES) model, and a group of Reynolds Averaged Navier- Incropera (1996) and is shown in Tabe 4. As for the nanofluid proper-
Stokes (RANS) equations (e.g., the standard k-𝜀, realizable k-𝜀, Spalart- ties, the copper oxide nanofluid was introduced into the software as a
Allmaras and the standard k-𝜔 models) (Rehman, 2011). Both models homogeneous fluid by defining its effective parameters. The solid CuO
are used in predicting the turbulent/eddy viscosity which completes the nanoparticles are assumed to have an average size of 29 nm, a density
flow equations and can be later used to predict the eddy diffusivity for of 6500 kg/m3 , and a specific heat capacity of 549.12 J/kg-K, as in the
modeling the transfer of momentum (Bakker, 2002). The time-averaged study of Popa et al. (2016) Eqs. (8)., (9), and (10) were used to calculate
case model that was used for the study is under the RANS equations. the effective density (𝜌ef f ), specific heat capacity (𝐶 𝑝ef f ), and nanofluid
The turbulence model and discretization scheme that were recom- viscosity (𝜇), respectively, at different loadings. In these equations, 𝜌𝑝
mended due to their high accuracy obtained from the previous study is the density of the copper oxide nanoparticles, ∅ is the volume fraction
of Ozden and Tari (2010) were chosen; these include the realizable k-𝜀 or particle loading, and the bulk viscosity and density of water were
model and first order discretization scheme. The form of the transport represented as μbf and ρbf . All correlations were taken from Azad and
equation of the realizable k-𝜀 model that approximates the dissipation Azad (2016) Table 4. shows the effective propeties of the base fluid
rate (𝜀) is given in Eq. (7) with its constants given by Table 3. (water) and the CuO-water nanofluid used in this present work.
( ) ( )
{( ) } ( ) ρef f = 𝜙 ρnp + (1 − 𝜙) ρbf (8)
δ ( ) δ μ δε ε
ρεuj μ+ t +ρC1 Sε + C1ε C3ε Gb
δxj δj σε̃ δxj k
[ ] 𝜙ρnp Cpnp + (1 − 𝜙)ρbf Cpbf
ε2 Cpef f = (9)
−C2 ρ √ + Cε (7) ρef f
𝑘 + εv
( )
μ = μbf 1 + 39.11𝜙 + 533.9𝜙2 (10)
For the discretization schemes, the standard scheme is selected for
pressure, while first order upwind scheme is selected for momentum, In order to properly determine the extent of heat transfer enhance-
turbulent kinetic energy and dissipation rate. The convergence crite- ment, a carefully-selected effective thermal conductivity model was uti-
ria that were built-in standards in ANSYS were used in the study. This lized in this study Eq. (11). is the ’present model’ for thermal conduc-
convergence criteria are based on residuals (in the order of 10−3 ). The tivity of nanofluids, as proposed by Sitprasert et al. (2009), which ac-
residual is considered as one of the most fundamental measures of an counts for the effect of the temperature dependency of the interfacial
iterative solution’s convergence because of its direct quantification of layer thickness (h) of the solid-liquid interface and temperature, where
the error in the solution of the system of equations. In a CFD analy- rnp is the radius of the nanoparticle, and the two functions are given as
sis, the residual measures the local imbalance of a conserved variable follows: β = 1 + ha and β1 = 1 + 2a h
Eqs. (12). and (13) give the expres-
in each control volume. In an iterative numerical solution, the residual sion for the thickness and thermal conductivity of the interfacial layer,
will never be exactly zero. However, the lower the residual value is, the respectively. For CuO-water nanofluid, the constant C is calculated us-
more numerically accurate the solution. ing Eq. (14) (Sitprasert, 2009) Table 5. presented the effective thermal
As for the nanofluid, single-phase flow was used to model the fluid conductivity, keff , of the base fluid (water) and the CuO-water nanofluid
to simplify the simulation process. The single-phase approach justifies used in this present work.
that since the nanoparticles are fluidized in the suspension due to their ( ) [ ] ( ) [ ( ) ]
ultra-fine dimensions, the particles and the base fluid have the same knp − klr 𝜙klr 2β1 3 − β3 + 1 + knp + 2klr β1 3 𝜙β3 klr − kbf + kbf
velocity and flow behavior.
kef f = ( ) ( ) [ ]
β1 3 knp + 2klr − knp − klr 𝜙 β1 3 + β3 − 1
Fluid properties. The thermal properties were introduced in the sys-
(11)
tem as temperature independent, except for the thermal conductivity.
The constant properties were taken from a reference temperature of
300 K (inlet temperature) while the temperature dependency of ther- h = 0.01(T − 273)rnp 0.35 (12)

4
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Table 5 Number (Re), are given in Table 7.


Effective thermal conductivity of base fluid (Water)a and CuO-water nanofluidb
Duρ
used. Re = (15)
μ
T keff (W/m-K)
(K) 2.3. Post-Processing
0.0 vol % 0.1 vol % 0.25 vol % 0.50 vol % 1 vol % 2 vol %

300 0.61 0.61 0.61 0.61 0.61 0.62


The post-processing stage regulates the desired output from the sim-
305 0.62 0.62 0.64 0.65 0.66 0.66
310 0.63 0.63 0.66 0.68 0.70 0.72 ulations. ANSYS provides automatic reports for the fluxes and surface
315 0.63 0.64 0.68 0.72 0.75 0.78 area variables after calculating the solutions. The total heat transfer rate,
320 0.64 0.65 0.70 0.75 0.80 0.85 shell outlet temperature, and pressure drop are the parameters that are
325 0.65 0.67 0.72 0.78 0.85 0.92 directly taken from the ANSYS Fluent results. Here, the heat transfer
330 0.65 0.68 0.74 0.81 0.90 0.99
335 0.66 0.70 0.77 0.85 0.95 1.06
coefficient, h, was calculated using Eq. (16).
340 0.66 0.72 0.80 0.89 1.00 1.13 Q
345 0.67 0.74 0.84 0.93 1.06 1.21 h= ( ) (16)
350 0.67 0.77 0.88 0.99 1.12 1.29
A Twall − Tbulk

a
Properties taken from Incropera (1996). A = nπDL (17)
b
Properties calculated using Eqs. (11) to (14).
Where Q is the total heat transfer rate, A is the heat transfer area which
Table 6 is determined using Eq. (17), Twall is the average wall temperature, Tbulk
Boundary conditions for the investigated system. is the bulk temperature and is the average of the inlet and outlet shell
temperature, D is the diameter of the pipe, and L is the length of the
STHE operating conditions Values
pipe.
Gauge pressure (outlet nozzle) 0 The performance index (η) was evaluated from parameters that were
Shell inlet temperature (K) 300 obtained from the post-processing stage as given by Eq. (18). It is defined
Wall temperature (K) 450
by Tabari et al (2016) as the ratio of the heat transfer and pressure
Mass flow rate (inlet nozzle) (kg/s) 0.5, 1, 2
drop improvement. The higher the performance index, the higher is the
potential practical use of the characterized parameters for the nanofluid.
Table 7 The performance index will also be the main basis for the determination
Velocities (m/s) of varying concentration of CuO-water nanofluid at differ- of the extent of the effects of the particle loading and flow rate.
ent Reynolds number.
hnf
Loading (%vol) Re = 17,804 Re = 35,559 Re = 71,121 hb REh
η= ΔPnf
= (18)
REΔP
0.1 0.51 1.02 2.05 ΔPb
0.25 0.54 1.08 2.16
0.5 0.58 1.17 2.34 In addition to these post-processing parameters, temperature con-
1 0.68 1.36 2.72 tours and flow profiles are also determined to aid in the assessment of
the performance of nanofluid as a coolant in the shell and tube heat
exchanger.

Chkbf 3. Result and discussion


klr = (13)
rnp
3.1. Validation of the CFD model
0.1
C= (T − 298) (14) An effective model should agree within a reasonable range with ei-
(Φ + 0.001)
ther the experimental data, theoretical values, or a previously validated
Boundary conditions. The boundary conditions used by Ozden and CFD modelling approach depending on the complexity of the geome-
Tari (2010), as shown in Table 6, were assigned for the current model. try and system to be defined. In this study, the basis of verification was
No-slip boundary condition is assumed for all surfaces, whereas the shell based from the last criteria where the current runs were compared to the
side is assumed to be perfectly insulated by assigning zero heat flux. CFD results of Ozden and Tari (2010), which were experimentally and
Velocity inlet was chosen as the type of boundary input for the inlet theoretically validated from their previous studies. The results from the
nozzle in the nanofluid runs to maintain the same Reynolds number for current CFD model, together with the corresponding differences with
a given mass flow rate run. the data of Ozden and Tari (2010) are summarized in Table 2. In gen-
Model validation and parameter variation. The model was validated by eral, the differences increase with increasing mass flow rate. The cur-
conducting preliminary simulations using water for the shell fluid. The rent model over-predicted the temperature of the shell outlet by as low
results of these simulations were then compared to the similar set-up of as 0.28 K, with 2.2 K being the highest. It consistently overshoots the
the experimentally-verified CFD results of Ozden and Tari (2010). The values for the total heat transfer rate by 0.4 to 6.40%, while the pressure
CuO-water nanofluid was introduced as the shell fluid once the results drop results were under-predicted by approximately 20%. The high er-
of the ANSYS Fluent analysis were agreeable to the mentioned runs. ror from the pressure drop can be justified by looking at the analytically
The study was conducted at a fixed Reynolds number, an effective & calculated values of the pressure drop from Ozden and Tari’s previous
common method of assessing the effect of varying the concentration of work, as summarized in Table 8. Based from this, the actual difference
nanofluids. The volume fraction of the CuO-water nanofluid was varied only lies at around 8%, with the lowest mass flow rate condition devi-
(0.1, 0.25, 0.50, and 1%vol) and were individually subjected to differ- ating by only 0.85%. The bulk of the error could possibly come from
ent values of Reynolds number, which correspond to the different mass the differences in the meshing and minor details from the geometry and
flow rates based on the study of Ozden and Tari (2010). The Reynolds modeling that were not specified. The same trend of error can be ex-
numbers were calculated using Eq. (15), where the approaching inlet pected from the nanofluid runs but of slightly higher values as there are
velocity, u, in the nozzle was directly taken from ANSYS. The result- initially existing errors coming from the deviations of Ozden and Tari’s
ing inlet velocities, for varying nanofluid particle loading and Reynolds model from the experimental and theoretical values.

5
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Table 8
Comparison of the pressure drop (Pa) to theoretical and simulation values.

Mass flow rate (kg/s) This work CFD analysis Percentage Bell-Delaware calculation Percentage
(Ozden and Tari, 2010) difference (%) (Ozden and Tari, 2010) difference (%)

0.5 1,231.5 1,509 20.25 1,242 0.85


1 4,964.53 6,112 20.72 4,592 7.80
2 20,164.2 24,464 19.27 18,630 7.91

Fig. 4. Heat transfer coefficient at varying concentration of CuO-water nanofluid and Reynolds number–○, Water; X, 0.1%vol; □, 0.25%vol; ∗ , 0.5%vol; ◇, 1%vol;
and lines, guide for the eyes.

Table 9 tive at lower turbulence regions. The fluid at this region is flowing at
Percentage increase of heat transfer coefficient at varying Reynolds number a lower velocity, thus having a larger residence time along the length
(Re). of the heat exchanger, allowing a more effective contact for the heat
Loading (%vol) Re = 17,804 Re = 35,559 Re = 71,121 transfer. Therefore, it is significant to keep the fluid flow of nanofluids
at lower turbulence regions upon its application as coolant. This obser-
0.1 12.19 10.75 9.76
0.25 22.26 19.61 18.20
vation is also made by Pantzali et al. (2009), in their study of nanofluids
0.50 33.13 29.73 27.86 in a brazed plate heat exchanger.
1 48.16 43.83 41.67

3.3. Pressure drop of CuO-water nanofluid


3.2. Heat transfer of CuO-water nanofluid
The main penalty of the use of higher concentrations of nanofluid is
Fig. 4 shows the effect of implementing different concentrations of the resulting increase in the pressure drop. Fig. 5 shows the magnitude
nanofluid on its heat transfer characteristic. It shows that increasing the of these consequences at varying concentrations and Re. A clear trend
concentration of the nanofluid drastically enhances the heat transfer can be seen in the figure, wherein increasing the particle loading and
coefficient. The said observation is primarily due to the very strong in- Re significantly increases the pressure drop. The increase is contributed
crease in the effective thermal conductivity of the fluid. A particle load- by the higher velocities needed for higher particle loading to maintain a
ing as low as 0.1%vol improves the heat transfer coefficient to about 9 constant Re. Another main factor for the said observation is the higher
to 12%, and can go as high as 41 to 48% enhancement for the highest viscosity of nanofluids at higher particle loading.
particle loading, depending on the Reynolds Number (Re). Notably, for all cases of Re, at the highest loading (1%vol) the pres-
The increase in the convective heat transfer coefficient of nanofluid sure drop approximately doubles, while at the lowest concentrations,
flows are further augmented with increasing the Re. However, higher the pressure drop is found to increase by only 6 to 7%. It is worth men-
Re does not guarantee that the effectiveness of the improvement; that tioning that the increase of the pressure drops at increasing Reynolds
is, the percentage increase in heat transfer rate or heat transfer coef- number is non-linear, compared to that of the almost linear trend in
ficient relative to the base fluid is larger. In fact, the difference of the the increase found in the heat transfer rate. The results also show that
percentage enhancement between the lowest and highest Re is around for particle loading greater than 0.25%vol, the percentage increase in
6.5% for the nanofluids with particle loading of 1%vol. A summary of the pressure drop is always higher than the increase in the heat transfer
the observations can be seen in Table 9, in which the said behavior is rate, regardless of the Reynolds number. The nature of the two curves
true for all particle loading of the nanofluid. It is also to be noted that (heat transfer coefficient and pressure drop versus Re) also emphasize
the difference in the percentage increase of the heat transfer coefficient this observation, in which the differences when the concentrations are
at two different Reynolds number increases as the particle loading in- varied are more pronounced. These statements can better be visualized
creases. The results verify the fact that nanofluids are much more effec- and analyzed by looking at the performance index.

6
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Fig. 5. Pressure drop at varying concentration of CuO-water nanofluid and Reynolds number–○, Water; X, 0.1%vol; □, 0.25%vol; ∗ , 0.5%vol; ◇, 1%vol; and lines,
guide for the eyes.

Fig. 6. Performance indices of CuO-water nanofluid at different Reynolds number and particle loading–□, 0.1%vol; X, 0.25%vol; ○, 0.5%vol; and ∗ , 1%vol.

3.4. Performance index that the enhancement is almost still being off-set by the pressure drop
penalty.
The performance index is a great indicator of the overall effective- At the highest concentration, 1%vol, the pressure drop increase is
ness of the nanofluid flow as it takes in account the enhancements for very significant as compared to the heat transfer enhancements, regard-
both its desirable (heat transfer) and undesirable (pressure drop) effects. less of the Reynolds number, and therefore is not advisable for future
The extent at which one limits the other can only be emphasized by de- applications. This observation is consistent with the findings recorded
termining the performance index. in literature, in which the best concentration for nanofluid use is found
Fig. 6 presents the performance indices of nanofluid flow at vary- to be lower than 1%vol.
ing conditions. It shows that flow conditions that have higher particle
loading and Reynolds number results to a relatively lower performance
3.5. Analysis of CuO-water nanofluid behavior in the STHE
index. Moreover, it was found that the only nanofluid conditions that
are practical for application in the shell and tube heat exchanger are the
Determination of the performance index for the fluid conditions used
0.1%vol particle loading under all Reynolds numbers, and the 0.25%vol
in the CFD analyses is important to establish the usefulness of CuO-
of the first two regimes, as they produced indices greater than 1. Despite
water nanofluid as coolant in the shell and tube heat exchanger (STHE).
this, the magnitude of the indices (being very close to unity) signifies
Based on the heat transfer and pressure drop studies conducted, the

7
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Fig. 7. Comparison of the cross-sectional temperature contours for (A) water; and (B) 0.1%vol CuO-water nanofluid (right) at Re = 17,804.

nanofluid that produced the highest performance index is under the and the fluid near the inner tube are almost equal in temperature
conditions of low Reynolds number (Re = 17,804), and low particle Rehman (2011)., in his CFD analysis of an unbaffled shell and tube
loading (0.1%vol). However, nanofluids are commonly used in various heat exchanger, found that the inner fluid that is closer to the tubes
equipment such as radiators and plate heat exchangers but seldom in contributes more to heat transfer than the outer fluid. It was concluded
shell and tube heat exchangers. The proceeding sections then highlight that this is due to fluid velocity and heat exchanger geometry. Thus, the
the performance of 0.1%vol CuO-water nanofluid in the shell and tube observations found in the current model may also be attributed to the
heat exchanger. tube pitch, baffles, and shell configuration used in the model.
Figs. 8 and 9 show the overall temperature profile of the cross-
3.6. Temperature distribution analysis sections previously shown in the thermal contours. The curves display
the thermal behavior of the fluids with respect to the shell radius. Inter-
Temperature contours were generated along the length of the heat estingly, similar to the study of Rehman (2011), in all cases, it is evident
exchanger and on different cross-sectional surfaces to emphasize the that the fluid near the inner tube is more heated than the fluid near the
temperature distribution for the steady-state flow of water and nanofluid outer tubes. Thus, most of the heat transfer occurs due to the fluid near
as coolants. Fig. 7 shows the differences between the cross-sectional tem- the inner tube, as depicted by the high temperature gradient near the
perature contours of water and 0.1%vol CuO-water nanofluid. The posi- center of the figures, versus the lower temperature gradient of the fluid
tions selected display the thermal behavior of the fluid in the inlet and far from the center of the shell. In the figures, it can be seen that re-
outlet regions, and also on surfaces before, after and in between the baf- gardless of the fluid used, heating is uniform until approximately 0.30
fles. As expected, the fluid temperature is uniform at the inlet region. m of the heat exchanger, wherein a temperature gradient begins to ap-
Also, for both the nanofluid and water, a temperature gradient ranging pear, as previously observed. Although there is no immediate difference
from 360 to 400 K is evident near the tube walls. However, a notable if the temperature contours of water and nanofluid are compared, the
difference in temperature initially became evident in the cross-sections bulk temperature of the nanofluid is actually higher than the bulk tem-
of both water and the nanofluid at about 0.30 m of the heat exchanger, perature of water, with values equal to 368.3 and 366.4 K, respectively.
or the area before the second baffle. At 0.45 m, this difference is more Finally, despite the slight difference in the average bulk tempera-
pronounced, such that the upper cross-sectional portion of the shell is ture of the fluids, this allowed better heat transfer of the shell fluid with
darker green in color (with temperature range of 360 to 395 K) for the the tubes, until the outlet temperature is reached. Together with the
nanofluid, compared to the thermal contour of water at the same posi- lower heat capacity compared to that of water, the increase in bulk
tion. This observation can be confirmed by simply comparing the spe- temperature may also be attributed to the higher thermal conductiv-
cific heat capacity of the fluids used. For example, at 300 K, water has a ity of the nanofluid, and this is also observed by Anbumeenakshi and
heat capacity of 4182 kJ/kg-K, while the 0.1%vol CuO-water nanofluid Thansekhar (2017) in their study of a microchannel heat sink cooled
has a lower heat capacity that is equal to 4158 kJ/kg-K. Although this using a nanofluid. Note that the discussions made simply provide sup-
difference is minimal, the lower heat capacity of the nanofluid allows porting details for the observations previously made in the thermal be-
it to absorb more heat easily than water, thus producing a higher bulk havior of the bulk fluids based on the temperature contours.
fluid temperature, as seen in the thermal contours. This observation is Fig. 10 shows the temperature contour along the symmetrical plane
also made by Bhanuteja and Azad (2013), in which they studied various of the heat exchanger using 0.1%vol CuO-water nanofluid. As previ-
nanofluids in shell and tube heat exchanger using CFD analysis. ously noted, at about 0.45 meters, the upper part of the shell is darker
Aside from the positions mentioned, it is also remarkable that for in color than the rest of the areas that is not affected by the fluid flow.
the other cross-sections shown in Fig. 7, the fluid near the outer tubes Thus, in order to further understand the thermal behavior of the fluid

8
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Fig. 8. Cross-sectional temperature profile for water at Reynolds number = 17,804.

Fig. 9. Cross-sectional temperature profile for CuO-water nanofluid (0.1%vol) at Reynolds number = 17,804.

Fig. 10. Temperature contour along the length of the heat exchanger for CuO-Water nanofluid (0.1%vol and 1%vol).

9
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Fig. 11. Average temperature profiles along the length of the heat exchanger for CuO-water nanofluid (0.1%vol) and water at three Reynolds numbers.

with respect to the length of the heat exchanger, an average tempera- Table 10
ture profile is generated at three different Reynolds number, using the Effect of pressure drop on outlet temperature at varying loading
particle loading that produced the highest performance index (Fig. 11). and Re.
For guidance, Reynolds numbers 1, 2, and 3, have a value of 17804, A. Re = 17,804
35,559, and 71,120, respectively.
Loading (%vol) Pressure Drop (Pa) Outlet Temperature (K)
As expected, the temperature gradient produced by the 0.1%vol
nanofluid is the highest, and this can be one of the reasons why it pro- 0.1 1,313.84 337.03
0.25 1,457.96 338.23
duced the highest performance index as well. It can also be observed
0.50 1,734.90 338.39
that increasing the Reynolds number (which causes an increase in inlet 1 2,413.22 337.01
fluid velocity), produced lower outlet temperatures. This is consistent Mean 1,729.98 337.67
with the observations of Ozden and Tari (2010), in which increasing
B. Re = 35,559
the mass flow rate of the shell fluid decreases the outlet temperature.
For a given Reynolds number, the thermal profiles of the 0.1%vol Loading (%vol) Pressure Drop (Pa) Outlet Temperature (K)
nanofluids produced a higher temperature gradient and consequently, 0.1 5,319.07 331.66
a higher outlet temperature than their water counterpart. This testifies 0.25 5,901.65 332.46
the enhancements in heat transfer due to nanofluids. In fact, if a straight 0.50 7,026.38 332.47
1 9,757.81 331.18
line is drawn between the temperature profiles of water and nanofluid
Mean 7,001.23 331.94
(Re 1, for example), it is evident that the outlet temperature (attained at
about 0.57 m) of water is already attained by the nanofluid at only about C. Re = 71,120
0.40 m of the heat exchanger. However, it is interesting to note that the Loading (%vol) Pressure Drop (Pa) Outlet Temperature (K)
thermal behavior of the fluids is the same with respect to the heating
0.1 21,502.73 329.56
area of the equipment. That is, in all cases, heat transfer occurred mostly
0.25 23,919.57 330.22
at the region between the inlet up to approximately 0.13 m of the heat 0.50 28,482.74 330.17
exchanger, as represented by the steep temperature gradient in all the 1 39,655.74 328.93
curves shown in Fig. 11. The fluid continues to heat up until about 80% Mean 28,390.20 329.72
of the length of the heat exchanger, where the temperature gradient
starts to decrease up to the outlet nozzle. According to Rehman (2011), for shell and tube heat exchangers, the
comparable pressure gradient observed around the inlet and outlet re-
3.7. Pressure contours and profiles gions may be due to the fluid impingement at the heat exchanger noz-
zles. Additionally, Teng et al. (2011), in their study of TiO2 nanofluid
Fig. 12 depicts the pressure drop as a function of heat exchanger in circular pipes, concluded that the pressure drop in circular pipes de-
length, for all four particle loadings tested under the lowest Reynolds creases with an increase in temperature. The present results agrees to
number (Re = 17,804). In all cases, the highest-pressure gradient was the previous works, that is, at constant Reynolds number and increasing
observed along the length of the heat exchanger, which may be caused particle loading up to below 1%vol, the increase in shell side pressure
by the baffles present inside the shell, and also by the tube configuration. drop causes a decrease in the outlet temperature Table 10. presented
The pressure difference present at the inlet and outlet nozzle regions are these results.
also notable, although the pressure at the inlet is almost half of the pres- For a constant Reynolds number, the profiles shown in Fig. 12 also
sure drop present within majority of the length of the heat exchanger. represent the increase in pressure drop with increasing particle loading.

10
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Fig. 12. Pressure profile along the length of the heat exchanger for CuO-water nanofluid.

Fig. 13. Pressure contour along the length of the heat exchanger for CuO-Water nanofluid (0.1%vol) at Reynolds number = 17,804.

This trend is mainly due to the increase in viscosity of the nanofluid fluid flow inside the shell, thereby maximizing the heat transfer area.
with particle loading. However, unlike the temperature profiles which For any Reynolds number, a comparison made between the velocity con-
showed that the thermal behavior of the fluids is similar, regardless of tours of the nanofluids versus that of water showed no significant differ-
particle loading, the pressure profile showed different pressure gradi- ences, except for the magnitude of the velocities. The similarities found
ents that are characteristic of the fluid properties. It is indeed impor- in flow behavior is somewhat expected because the nanofluids in the
tant that nanofluid coolants with higher particle loading be operated at current study are modeled as single-phase fluids, and thus only the best
high velocities, in order to overcome the increased resistance due to the nanofluid condition is displayed in the figure. Additionally, based on
viscosity increase. Doing so may also alleviate the effects of pressure the heat transfer study conducted, for a constant Reynolds number, the
drop on the overall performance index of the nanofluid. Fig. 13 simply outlet temperature increases only for particle loading below 1%vol. For
shows the differences in pressure along the length of heat exchanger particle loading equal to 1%vol, a slight decrease in the outlet tempera-
using 0.1%vol CuO-water nanofluid. ture of the fluids is observed, and this may be due to the 14% difference
in the velocity used for 1%vol nanofluid versus the velocities of those
3.8. Fluid flow behavior and vector plot analysis below 1%vol. This however, does not affect the average bulk tempera-
ture of the nanofluids, since the heat capacity and thermal conductivity
Fluid flow along the whole heat exchanger is shown as a contour of the nanofluids are more desirable than that of water.
in Fig. 14, as represented by the yellow and green regions (with ve- Similar to the thermal profiles, the cross-sectional velocity profile
locity ranging from 0.30 to 0.55 m/s). The velocity contour, together shown in Fig. 15 depicts the flow distribution in each cross-section along
with the temperature and pressure contours that are previously shown the heat exchanger, for 0.1%vol CuO-water nanofluid. Due to the shell-
as a function of heat exchanger length (Figs. 10 and 13, respectively), side configuration and presence of baffles, the velocity profile for the
provide further understanding of the heat transfer and fluid flow along cross-sections of the heat exchanger is not developed. Interestingly, the
the heat exchanger. As expected, it can be seen that the baffles direct profiles for both water and nanofluid, are almost similar, in which the

11
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Fig. 14. Velocity contour along the length of the heat exchanger for CuO-water nanofluid (0.1%vol) at Reynolds number = 17,804.

Fig. 15. Cross-sectional velocity profile for CuO-water nanofluid (0.1%vol).

fluid near the outer tubes produced a higher velocity gradient than the
fluid near the center of the shell. Regardless of the particle loading, the
fluid near the center contributes more to heat transfer due to lower local
velocities and higher temperature gradients. Thus, similar to the obser-
vations made by Rehman (2011), lower fluid velocity near the center
results to higher residence time, which then maximizes heat transfer.
This is also previously observed in the cross-sectional thermal profiles.
However, more peaks are observed in the nanofluid velocity profiles,
which may be due to the higher inlet velocities used.
Vector plots and streamlines are also generated in order to further
understand the effects of nozzle and tube configuration on the fluid pres-
sure drop. The vector plots at the inlet and outlet regions for the 0.1%vol
nanofluid are shown in Fig. 16, and since it is previously observed that
there are no major variations between the flow behavior of water and
nanofluid, only one set of vector plot is displayed. According to a shell
and tube heat exchanger design proposed by Al-Hadhrami (2009), fluid
recirculation causes stagnation of fluid behind the baffles, which con-
sequently lead to fouling and corrosion of the tube walls. Thus, it is
important to identify and minimize the recirculation zones to achieve
an effective design. For the current model, an observation of the vector
plots for all conditions show that most of the recirculation zones can
only be found at the inlet. The swirls represent the recirculation at the Fig. 16. Vector plots at the inlet (A) and outlet (B) regions for 0.1%vol CuO-
top, center, and bottom areas of the inlet region. water nanofluid.

12
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

Fig. 17. Streamlines that emphasize the recirculation zones using (A) 0.1%vol CuO-water nanofluid flow; and (B) water is used as shell fluid.

Interestingly, as the flow expands from the nozzle to the shell, it Performance indices in general were high at conditions of low par-
is possible that a boundary layer separation occurs. This internal flow ticle loading and low Reynolds number. These conditions produced in-
separation may have caused formulation of the recirculation zones due dices that are greater than 1, namely nanofluids operated at 0.1 and
to the high-pressure gradient that is present at the inlet nozzle region 0.25%vol, for Reynolds number equal to 17,804 and 35,559. Thus, par-
(Wilcox, 2007). This observation seems to be consistent for the zero- ticle loadings greater than 1%vol are not recommended for shell and
pressure outlet nozzle, in which no recirculation zones were found. tube heat exchanger application since the penalty in pressure drop is
Fig. 17 shows the streamlines for CuO-water nanofluid and water at the significantly high.
lowest Reynolds number. Interestingly, despite the increase in nanofluid The temperature profiles generated provided information on the ef-
viscosity due to the increase in particle loading, no recirculation zones fects of the nanofluid properties to its thermal behavior. A more uniform
are found in the heat exchanger (except for those found at the inlet), if heating is observed for the nanofluid based on the cross-sectional con-
0.1%vol CuO-water nanofluid is used as shell fluid. Note that there are at tours generated. Meanwhile, from the cross-sectional temperature pro-
least two other recirculation zones found in the streamlines for water as files, the average bulk temperature of the nanofluid is higher than that
shell fluid. Although not shown, the same recirculation zones were also of water because of its lower heat capacity, which allows for more heat
evident in the streamlines for 1%vol nanofluid under the same Reynolds to be absorbed. This is considered advantageous because as heat transfer
number. Although these observations add to the benefits of utilizing the progresses, the bulk temperature will continually increase until an outlet
lowest particle loading at the lowest fluid velocity, the shell and tube temperature, that is higher than that of water, is produced. This is fur-
heat exchanger studies of both Al-Hadhrami (2009), and Ozden and ther observed when the temperature profiles along the length of the heat
Tari (2010) claim that the development of a recirculation zone is mainly exchanger, for three different Reynolds numbers and 0.1%vol particle
affected to the design of the heat exchanger, particularly by the baffle loading, are compared. It was found that all the nanofluids produced
cut and baffle spacing used. an outlet temperature that is higher than water, only if the Reynolds
number used is low. In general, due to the similarities in the profiles,
the particular thermal behavior of both nanofluid and water in the shell
4. Conclusion and tube heat exchanger are similar, such that the highest temperature
gradient occurs at approximately 0.13 meters of the heat exchanger. Al-
The CFD analyses of the nanofluids showed that increasing the load- though this is the case, the enhancements in outlet temperature, and
ing of CuO-water nanofluid results to an increase in both heat transfer consequently in heat transfer, are more evident for the nanofluid.
and pressure drop. However, the heat transfer enhancement was more The pressure profiles and contours generated also aided the identifi-
pronounced at lower turbulence regions. Introduction of nanoparticles cation of the regions where pressure drop occurs. It was found that most
in the base fluid as low as 0.1%vol can increase the heat transfer by of the pressure drop is induced by the configuration and number of baf-
9 to 12%, with the highest particle loading (1%vol) producing 41-48% fles installed in the shell, although the pressure drop at the inlet region
enhancement. Consequently, higher loading produces higher pressure is also notable. For a constant Reynolds number, the pressure drop in
drop. The percentage increase in the pressure drop is non-linear as com- the heat exchanger increases with increasing particle loading. Thus, it
pared to the heat transfer rate with increasing Reynolds number. Re- is important that nanofluids used at a high particle loading be operated
gardless of the Reynolds number, the particle loading of 1%vol produces at high velocities to overcome the effects of increased viscosity.
a pressure drop that is approximately double in magnitude compared to An observation of the flow behavior of the nanofluid showed that
the pressure drop due to water. The percentage increase in the pressure due to the single-phase modeling employed, the nanofluid flow behaves
drop is much larger versus the increase in heat transfer rate at concen- similarly to water. It was also found that nanofluids operated at a par-
trations greater than 0.25%vol. ticle loading of 1%vol, regardless of the Reynolds number, produce a

13
P.A.D. Cruz, E.E. Yamat, J.P.E. Nuqui et al. Digital Chemical Engineering 3 (2022) 100014

lower outlet temperature. This may be attributed to the increase in ve- ANSYS Inc. (2013). ANSYS CFX Reference Guide. 15th ed., ANSYS, Inc.
locity which was necessary to overcome the effects of pressure drop. Azad, A., Azad, N.V., 2016. Application of nanofluids for the optimal design of shell and
tube heat exchangers using genetic algortihm. Case Stud. Therm. Eng. 8, 198–206.
Finally, an illustration of the vector plots and streamlines show that due Bakker, A. (2002). Lecture 10 - Turbulence models–Applied Computational Fluid Dynam-
to the contraction in flow due to the nozzle, a high-pressure gradient is ics.
observed at the inlet region that produced recirculation zones. Bhutaneja, S., Azad, D., 2013. Thermal performance and flow analysis of nanofluids in a
shell and tube heat exchanger. Int. J. Mech. Eng. Tech. 4, 164–172.
Overall, based on the single-phase modeling conducted, the appli- Bhutta, M.M.A., Hayat, N., Bashir, M.H., Khan, A.R., Ahmad, L.N., Khan, S., 2011. CFD
cation of CuO-water nanofluid in a shell and tube heat exchanger is applications in various heat exchangers design–A review. Appl. Therm. Eng. 32, 12.
determined by several parameters which include nanofluid properties, Devi, K.M., Nagamani, G.V., 2015. Design and thermal analysis of shell and tube
heat exchanger by using fluent tool. Int. J. Mag. Eng. Technol. Manag. Res. 2,
particle loading, and fluid velocity. Since it is evident that nanofluids
359–366.
cause increased thermal performance, the effects of pressure drop can Haghighi, E.B., 2015. Single Phase Convective Heat Transfer with Nanofluids–An Experi-
be overcome, and the heat transfer enhancement can be maximized, if mental Approach. Royal Institute of Technology, Stockholm, Sweden Doctoral Thesis.
Jadhav, A.D., Koli, T.A., 2014. CFD Analysis of Shell and Tube Heat Exchanger to study
the CuO-water nanofluid is operated at a low Reynolds number and low
the effect of baffle cut on the pressure drop. Int. J. Res. Aeronaut. Mech. Eng. 2,
particle loading in the shell and tube heat exchanger. 1–7.
Jokar, A., O’Halloran, S.P., 2013. Heat transfer and fluid flow analysis of nanofluids in
Funding corrugated plate heat exchangers using computational fluid dynamics simulation. J.
Therm. Sci. Eng. Appl. 5, 1–10.
Masilungan-Manuel, J.T., Manuel, M.C., Lin, P.T., Soriano, A.N., 2015. Optimization of
This research did not receive any specific grant from funding agen- the drying parameters for the short-form spray dryer producing powdered egg with
cies in the public, commercial, or not-for-profit sectors. 20% tapioca starch additive. Adv. Mech. Eng. 7, 1–11.
Othman, K.H., 2009. CFD Simulation of Heat Transfer in Shell and Tube Heat Exchanger.
Univeristi Malaysia Pahang, Malaysia Undergraduate Thesis.
Declaration of Competing Interest Ozden, E., Tari, I., 2010. Shell side analysis of a small shell-and-tube heat exchanger.
Energy Convers. Manag. 51, 1004–1014.
Pantzali, M.N., Mouza, A.A, Paras, S.V., 2009. Investigating the efficacy of nanofluids as
The authors declare that they have no known competing financial coolants in plate heat exchangers (PHE). Chem. Eng. Sci. 64, 3290–3300.
interests or personal relationships that could have appeared to influence Popa, C.V., Nguyen, C.T., Gherasim, I., 2016. New specific heat data for Al2 O3 and CuO
the work reported in this paper. nanoparticles in suspension in water and ethylene glycol. Int. J. Therm. Sci. 111,
108–115.
Puliti, G.S.Paolucci, Sen, M., 2011. Nanofluids and Their Properties. Appl. Mech. Rev. 64.
CRediT authorship contribution statement Rehman, U.U., 2011. Heat Transfer Optimization of Shell-and-Tube Heat Exchanger
through CFD Studies. Chalmers University of Technology, Goteborg, Sweden Master’s
Thesis.
Patricia Anne D. Cruz: Software. Ed-Jefferson E. Yamat: Soft-
Sitprasert, C., Dechaumphai, P., Juntasaro, V., 2009. A thermal conductivity model
ware. Jesus Patrick E. Nuqui: Writing – review & editing. Allan N. for nanofluids including effect of the temperature-dependent interfacial layer. J.
Soriano: Formal analysis. Nanopart. Res. 11, 1465–1476.
Skočilas, J., Palaziuk, I., 2015. CFD simulation of the heat transfer process in a chevron
References plate heat exchanger using the SST turbulence model. J. Adv. Eng. 55, 267–274.
Tabari, Z.T., Heris, S.Z., Moradi, M., Kahani, M., 2016. The study on application of
TiO2 /water nanofluid in plate heat exchanger of milk pasteurization industries. Re-
Al-Hadhrami, L.M. (2009). Shell and tube heat exchanger (US20090000775 A1). new. Sustain. Energy Rev. 58, 1318–1326.
Anbumeenakshi, C., Thansekhar, M.R., 2017. On the effectiveness of a nanofluid cooled Teng, T.P., Hung, Y., Jwo, C., Chen, C., Jeng, L.Y., 2011. Pressure drop of TiO2 nanofluid
microchannel heat sink under non-uniform heating condition. Appl. Therm. Eng. 113, in circular pipes. Particuology 9, 486–491.
1437–1443. Tiwari, A.K., Ghosh, P., Sarkar, J., 2015. Particle concentration levels of various nanofluids
Andersson, B., Andersson, R., Håkansson, L., Mortensen, M., Sudiyo, R., van Wachem, B., in plate heat exchanger for best performance. Int. J. Heat Mass Transf. 89, 1110–1118.
2012. Computational Fluid Dynamics for Engineers, 1st ed. Cambridge University Wilcox, David C., 2007. Basic Fluid Mechanics, 1st ed. DCW Industries, Inc.
Press.

14

You might also like